f\ ^™ F^ A
> C PA
      United States
          Air Quality Criteria for
          Ozone and Related
          Photochemical Oxidants
          (First External Review Draft)
          Volume III of

-------
                                                 EPA/600/R-05/004CA
                                                      January 2005
Air Quality Criteria for Ozone and  Related
           Photochemical  Oxidants
                    Volume
         National Center for Environmental Assessment-RTF Office
                Office of Research and Development
               U.S. Environmental Protection Agency
                  Research Triangle Park, NC

-------
                                   DISCLAIMER

     This document is an external review draft for review purposes only and does not
constitute U.S. Environmental Protection Agency policy. Mention of trade names or
commercial products does not constitute endorsement or recommendation for use.
                                     PREFACE

     National Ambient Air Quality Standards (NAAQS) are promulgated by the United States
Environmental Protection Agency (EPA) to meet requirements set forth in Sections 108 and 109
of the U.S. Clean Air Act (CAA). Sections 108 and 109 require the EPA Administrator (1) to
list widespread air pollutants that reasonably may be expected to endanger public health or
welfare; (2) to issue air quality criteria for them that assess the latest available scientific
information on nature and effects of ambient exposure to them; (3) to set "primary" NAAQS to
protect human health with adequate margin of safety and to set "secondary" NAAQS to protect
against welfare effects (e.g., effects on vegetation, ecosystems, visibility, climate, manmade
materials, etc); and (5) to periodically review and revise,  as appropriate, the criteria and NAAQS
for a given listed pollutant or class of pollutants.
     In 1971, the U.S. Environmental Protection Agency (EPA) promulgated National Ambient
Air Quality Standards (NAAQS) to protect the public  health and welfare from adverse effects of
photochemical oxidants. The EPA promulgates the NAAQS on the basis of scientific
information contained in air quality criteria issued under  Section 108 of the Clean Air Act.
Following the review of criteria as contained in the EPA  document, Air Quality Criteria for
Ozone and other Photochemical Oxidants published in 1978, the chemical designation of the
standards was changed from photochemical oxidants to ozone (O3) in 1979 and a 1-hour O3
NAAQS was set. The 1978 document focused primarily  on the scientific air quality criteria for
O3 and, to a lesser extent, on those for other photochemical oxidants such as hydrogen peroxide
and the peroxyacyl nitrates, as have subsequent revised versions of the ozone document.
     To meet Clean Air Act requirements noted above for periodic review of criteria and
NAAQS, the O3 criteria document, Air Quality Criteria for Ozone and Other Photochemical
                                         Ill-ii

-------
Oxidants, was next revised and then released in August 1986; and a supplement, Summary of
Selected New Information on Effects of Ozone on Health and Vegetation, was issued in January
1992. These documents were the basis for a March 1993 decision by EPA that revision of the
existing 1-h NAAQS for O3 was not appropriate at that time. That decision, however, did not
take into account some of the newer scientific data that became available after completion of the
1986 criteria document.  Such literature was assessed in the next periodic revision of the O3 air
quality criteria document, which was completed in 1996 and provided scientific bases supporting
the setting by EPA in 1997 of an 8-h O3 NAAQS that is currently in force together with the 1-h
O3 standard.
     The purpose of this revised air quality criteria document for O3 and related photochemical
oxidants is to critically evaluate and assess the latest scientific information published since that
assessed in the above 1996 Ozone Air Quality Criteria Document (O3 AQCD), with the main
focus being on pertinent new information useful in evaluating health and environmental effects
data associated with ambient air O3 exposures. However, some other scientific data are also
presented and evaluated in order to provide a better understanding of the nature,  sources,
distribution, measurement, and  concentrations of O3 and related photochemical oxidants and
their precursors in the environment. The document assesses pertinent literature available
through 2004.
     The present draft document (dated January 2005) is being released for public comment and
review by the Clean Air Scientific Advisory Committee (CASAC) to obtain comments on the
organization and structure of the document, the issues addressed, the approaches employed in
assessing and interpreting the newly available information on O3 exposures and effects, and the
key findings and conclusions arrived at as a consequence of this assessment.  Public comments
and recommendations will be taken into account making any appropriate further revisions to this
document for incorporation into a Second External Review Draft. That draft will be released for
further public comment and CASAC review before last revisions are made in response and
incorporated into a final version to be completed by early 2006. Evaluations contained in the
present document will be drawn on to provide inputs to associated PM Staff Paper analyses
prepared by EPA's Office of Air Quality Planning and Standards (OAQPS) to pose options for
consideration by the EPA Administrator with regard to proposal and,  ultimately, promulgation of
decisions on potential retention or revision, as appropriate, of the current O3 NAAQS.

-------
     Preparation of this document was coordinated by staff of EPA's National Center for
Environmental Assessment in Research Triangle Park (NCEA-RTP). NCEA-RTP scientific
staff, together with experts from other EPA/ORD laboratories and academia, contributed to
writing of document chapters.  Earlier drafts of document materials were reviewed by non-EPA
experts in peer consultation workshops held by EPA. The document describes the nature,
sources, distribution, measurement, and concentrations of O3  in outdoor (ambient) and indoor
environments.  It also evaluates the latest data on human exposures to ambient O3 and
consequent health effects in exposed human populations, to support decision making regarding
the primary, health-related O3 NAAQS. The document also evaluates ambient O3 environmental
effects on vegetation and ecosystems, man-made materials, and surface level solar UV radiation
flux and global climate change, to support decision making on secondary O3 NAAQS.
     NCEA acknowledges the valuable contributions provided by authors, contributors, and
reviewers and the diligence of its staff and contractors in the preparation of this draft document.
                                         Ill-iv

-------
           Air Quality Criteria for Ozone and Related
                   Photochemical Oxidants
                 (First External Review Draft)


                         VOLUME I


Executive Summary	E-l

1.   INTRODUCTION  	1-1

2.   PHYSICS AND CHEMISTRY OF OZONE IN THE ATMOSPHERE 	2-1

    CHAPTER 2 ANNEX (ATMOSPHERIC PHYSICS/CHEMISTRY)	AX2-1

3.   ENVIRONMENTAL CONCENTRATIONS, PATTERNS, AND
    EXPOSURE ESTIMATES	3-1

    CHAPTER 3 ANNEX (AIR QUALITY AND EXPOSURE) 	AX3-1
                            III-v

-------
           Air Quality Criteria for Ozone and Related
                   Photochemical Oxidants
                 (First External Review Draft)
                            (cont'd)

                         VOLUME II
4.   DOSIMETRY, SPECIES HOMOLOGY, SENSITIVITY, AND
    ANIMAL-TO-HUMAN EXTRAPOLATION	4-1

    CHAPTER 4 ANNEX (DOSIMETRY)	AX4-1

5.   TOXICOLOGICAL EFFECTS OF OZONE AND RELATED
    PHOTOCHEMICAL OXIDANTS IN LABORATORY ANIMALS
    AND IN VITRO TEST SYSTEMS  	5-1

    CHAPTER 5 ANNEX (ANIMAL TOXICOLOGY)	AX5-1

6.   CONTROLLED HUMAN EXPOSURE STUDIES OF OZONE AND
    RELATED PHOTOCHEMICAL OXIDANTS 	6-1

    CHAPTER 6 ANNEX (CONTROLLED HUMAN EXPOSURE)  	AX6-1

7.   EPIDEMIOLOGICAL STUDIES OF HUMAN HEALTH EFFECTS
    ASSOCIATED WITH AMBIENT OZONE EXPOSURE	7-1

    CHAPTER 7 ANNEX (EPIDEMIOLOGY)	AX7-1

8.   INTEGRATIVE SYNTHESIS	8-1
                            Ill-vi

-------
           Air Quality Criteria for Ozone and Related
                   Photochemical Oxidants
                 (First External Review Draft)
                           (cont'd)
                         VOLUME
9.   ENVIRONMENTAL EFFECTS: OZONE EFFECTS ON
    VEGETATION AND ECOSYSTEMS  	9-1

10.  TROPOSPHERIC OZONE EFFECTS ON UV-B FLUX AND
    CLIMATE CHANGE PROCESSES 	10-1

11.  EFFECT OF OZONE ON MAN-MADE MATERIALS	11-1

-------
                                    Table of Contents

                                                                                    Page

List of Tables	I-xiv
List of Figures 	I-xvi
Authors, Contributors, and Reviewers	I-xviii
U.S. Environmental Protection Agency Project Team for Development of Air
     Quality Criteria for Ozone and Related Photochemical Oxidants 	I-xix
U.S. Environmental Protection Agency Science Advisory Board (SAB)
     Staff Office Clean Air Scientific Advisory Committee (CASAC)
     Ozone Review Panel	  I-xx
Abbreviations and Acronyms	I-xxi

9.    ENVIRONMENTAL EFFECTS: OZONE EFFECTS ON VEGETATION
     AND ECOSYSTEMS	  9-1
     9.1      INTRODUCTION  	  9-1
     9.2      METHODOLOGIES USED IN VEGETATION RESEARCH	  9-2
             9.2.1     Introduction	  9-2
             9.2.2     Methods Involving Experimental Exposures to O3	  9-3
                      9.2.2.1    "Indoor", Controlled Environment, and
                                Greenhouse Chambers 	  9-3
                      9.2.2.2    Field Chambers	  9-5
                      9.2.2.3    Plume Systems  	  9-7
                      9.2.2.4    Comparative Studies	  9-10
                      9.2.2.5    Ozone Generation Systems	  9-11
                      9.2.2.6    Experimental Exposure Protocols	  9-11
             9.2.3     Methods Involving Exposures to O3 in Ambient Air  	  9-12
                      9.2.3.1    Air-exclusion Systems 	  9-13
                      9.2.3.2    Natural Gradients	  9-13
                      9.2.3.3    Use of Chemical Protectants	  9-14
                      9.2.3.4    Biomonitoring	  9-16
                      9.2.3.5    Calibrated Passive Monitors	  9-24
             9.2.4     Numerical/Statistical Methodologies  	  9-25
             9.2.5     Improved Methods for Defining Exposure	  9-27
     9.3      SPECIES RESPONSE/MODE-OF-ACTION 	  9-28
             9.3.1     Introduction	  9-28
             9.3.2     Mechanisms of Ozone-Induced Plant Alterations	  9-31
                      9.3.2.1    Changes in Metabolic Processes	  9-32
                      9.3.2.2    Modifications of Plant Physiological Processes	  9-34
             9.3.3     Ozone Uptake by Leaves	  9-35
                      9.3.3.1    Possible Reactions Within the Leaf	  9-43
                      9.3.3.2    Toxicants Within the Wall Space	  9-46
                      9.3.3.3    Products of Ozone  	  9-48
                      9.3.3.4    Antioxidants Within the Apoplastic Space	  9-55
             9.3.4     Wounding and Pathogen Attack	  9-64
                      9.3.4.1    Peroxidases	  9-67
                      9.3.4.2    Jasmonic Acid and Salicylic  Acid  	  9-69
                      9.3.4.3    Stress-Induced Alterations in Gene Expression	  9-69

-------
                               Table of Contents
                                     (cont'd)
                                                                                 Page
        9.3.5     Primary Assimilation by Photosynthesis	  9-73
                 9.3.5.1     Photooxidation: Light Reactions	  9-73
        9.3.6     Alteration of Rubisco by Ozone:  Dark Reactions  	  9-74
        9.3.7     Carbohydrate Transformations and Allocation	  9-77
                 9.3.7.1     Lipid Synthesis	  9-79
        9.3.8     Role of Age and Size Influencing Response to Ozone	  9-82
        9.3.9     Summary	  9-84
9.4     MODIFICATION OF FUNCTIONAL AND GROWTH RESPONSES	  9-85
        9.4.1     Introduction	  9-85
        9.4.2     Genetics	  9-88
                 9.4.2.1     Genetic Basis of O3 Sensitivity	  9-90
        9.4.3     Environmental Biological Factors	  9-91
                 9.4.3.1     Oxidant-Plant-Insect Interactions	  9-92
                 9.4.3.2     Oxidant-Plant-Pathogen Interactions	  9-95
                 9.4.3.3     Oxidant-Plant-Symbiont Interactions	 9-100
                 9.4.3.4     Oxidant-Plant-Plant Interactions:  Competition	 9-101
        9.4.4     Physical Factors	 9-104
                 9.4.4.1     Light	 9-104
                 9.4.4.2     Temperature  	 9-108
                 9.4.4.3     Humidity and Surface Wetness  	 9-111
                 9.4.4.4     Drought and Salinity	 9-113
        9.4.5     Nutritional Factors	 9-116
        9.4.6     Interactions with Other Pollutants	 9-118
                 9.4.6.1     Oxidant Mixtures  	 9-119
                 9.4.6.2     Sulfur Dioxide  	 9-119
                 9.4.6.3     Nitrogen Oxides, Nitric Acid Vapor, and Ammonia .... 9-120
                 9.4.6.4     Hydrogen Fluoride and Other Gaseous Pollutants	 9-123
                 9.4.6.5     Acid Deposition  	 9-123
                 9.4.6.6     Heavy Metals  	 9-125
                 9.4.6.7     Mixtures of Ozone with Two or More Pollutants  	 9-125
        9.4.7     Interactions with Agricultural  Chemicals	 9-125
        9.4.8     Factors Associated with Global Climate Change 	 9-126
                 9.4.8.1     Ozone-Carbon Dioxide-Temperature Interactions	 9-128
                 9.4.8.2     Ozone-UV-B Interactions	 9-151
                 9.4.8.3     Interactions of Ozone with Multiple Climate
                            Change Factors  	 9-153
        9.4.9     Summary - Environmental Factors	 9-153
9.5     EFFECTS-BASED AIR QUALITY EXPOSURE- AND
        DOSE-RESPONSE INDICES 	 9-157
        9.5.1     Introduction	 9-157
        9.5.2     Summary of Conclusions from the Previous Criteria Document .... 9-158
        9.5.3     Use of Exposure Indices to Establish Exposure-Response
                 Relationships	 9-161
                                      Ill-ix

-------
                               Table of Contents
                                    (cont'd)
                                                                                Page
        9.5.4     Identifying Exposure Components That Relate to
                 Vegetation Effects	  9-166
                 9.5.4.1    Role of Concentration	  9-166
                 9.5.4. 2     Role of Duration	  9-169
                 9.5.4.3     Patterns of Exposure	 9-169
                 9.5.4. 4     Frequency of Occurrence of Peak Concentrations	  9-172
                 9.5.4.5     Canopy Structure 	 9-173
                 9.5.4. 6     Site and Climate Factors	  9-174
                 9.5.4 . 7      Plant Defense Mechanism - Detoxification  	  9-174
        9.5.5     Ozone Uptake or Effective Dose as an Index	  9-174
                 9.5.5.1    Models of Stomatal Conductance	  9-175
                 9.5.5.2    Nonlinear Response and Developing Flux Indices	  9-178
                 9.5.5.3    Simulation Models	  9-179
        9.5.6     Summary	  9-179
9.6     OZONE EXPOSURE-PLANT RESPONSE RELATIONSHIPS  	  9-181
        9.6.1     Introduction	  9-181
        9.6.2     Summary of Key Findings/Conclusions from Previous
                 Criteria Documents	  9-182
        9.6.3     Ozone Indices and Ambient Exposure 	  9-193
        9.6.4     Effects of Ozone on Annual and Biennial Species	  9-212
                 9.6.4.1    Effects on Growth, Biomass, and Yield of
                           Individual Species	  9-212
                 9.6.4.2    Effects on Plant Quality 	  9-217
                 9.6.4.3    Effects on Foliar Symptoms	  9-219
                 9.6.4.4    Other Effects	  9-220
                 9.6.4.5    Scaling Experimental Data to Field Conditions	  9-221
                 9.6.4.6    European Critical Levels  	  9-224
                 9.6.4.7    Summary of Effects on Short-Lived Species	  9-227
        9.6.5     Effects of Ozone on Long-Lived (Perennial) Species	  9-229
                 9.6.5.1    Herbaceous Perennial Species 	  9-229
                 9.6.5.2    Deciduous Woody Species	  9-234
                 9.6.5.3    European Critical Levels  	  9-240
                 9.6.5.4    Summary of Effects on Deciduous Woody Species	  9-240
                 9.6.5.5    Evergreen Woody Species 	  9-241
                 9.6.5.6    Summary of Effects on Evergreen Woody Species	  9-244
                 9.6.5.7    Scaling Experimental Data to Mature Trees	  9-245
        9.6.6     Studies with the Chemical EDU	  9-249
        9.6.7     Summary	  9-251
9.7     EFFECTS OF OZONE EXPOSURE ON NATURAL ECOSYSTEMS  	  9-254
        9.7.1     Introduction	  9-254
        9.7.2     Case Studies  	  9-256
                 9.7.2.1    Valley of Mexico 	  9-256
                 9.7.2.2    San Bernardino Mountains	  9-263
                 9.7.2.3    Sierra Nevada Mountains	  9-268
                                     III-x

-------
                              Table of Contents
                                    (cont'd)
                                                                              Page
                 9.7.2.4    Appalachian Mountains  	  9-270
                 9.7.2.5    Plantago Studies in the United Kingdom	  9-271
                 9.7.2.6    Forest Health in the Carpathian Mountains 	  9-271
                 9.7.2.7    Field Exposure System (FACE), Rhinelander,
                           Wisconsin	  9-273
        9.7.3     Landscape Condition	  9-275
        9.7.4     Biotic Condition  	  9-278
                 9.7.4.1    Ecosystems and Communities 	  9-278
                 9.7.4.2    Species and Populations	  9-284
                 9.7.4.3    Organism Condition	  9-287
        9.7.5     Ecosystem, Chemical, and Physical Characteristics
                 (water, soil)	  9-295
                 9.7.5.1    Nutrient Concentrations, Trace Inorganic and
                           Organic Chemicals	  9-295
        9.7.6     Ecological Processes	  9-296
                 9.7.6.1    Energy Flow	  9-296
                 9.7.6.2    Material Flow	  9-298
        9.7.7     Hydrological and Geomorphological  	  9-301
        9.7.8     Natural Disturbance Regimes 	  9-301
        9.7.9     Scaling to Ecosystem Levels	  9-302
                 9.7.9.1    Scaling from Seedlings to Mature Trees	  9-303
                 9.7.9.2    Surveys, Growth Correlations and Stand-Level
                           Modeling	  9-305
        9.7.10    Summary of Ecological Effects  of Ozone Exposure on
                 Natural Ecosystems	  9-312
9.8      ECONOMIC EVALUATION OF OZONE EFFECTS ON AGRICULTURE,
        FORESTRY AND NATURAL ECOSYSTEMS	  9-315
        9.8.1     Introduction	  9-315
        9.8.2     The Measurement of Economic Information	  9-317
        9.8.3     Understanding of Air Pollutants Effects on the Economic
                 Valuation of Agriculture and Other Vegetation in the
                 1996 Criteria Document 	  9-318
                 9.8.3.1    Agriculture	  9-319
                 9.8.3.2    Forests (Tree Species) and Natural Ecosystems	  9-322
        9.8.4     Studies Since 1996 of Ozone Exposure Effects on the
                 Economic Value of Agriculture, Forests, and Ecosystems	  9-323
        9.8.5     Limitations of Scientific Studies and Economic Information	  9-325
        9.8.6     Conclusions	  9-328
9.9      SUMMARY AND CONCLUSIONS FOR VEGETATION AND
        ECOSYSTEM EFFECTS	  9-329
        9.9.1     Introduction	  9-329
        9.9.2     Methodology	  9-331
        9.9.3     Mode-of-Action	  9-332
        9.9.4     Modification of Growth Response  	  9-334
                                    Ill-xi

-------
                                   Table of Contents
                                        (cont'd)
                                                                                   Pas
            9.9.5      Exposure Indices	 9-337
            9.9.6      Ozone Exposure-Plant Response Relationships 	 9-340
            9.9.7      Ecosystem Effects	 9-343
            9.9.8      Economics	 9-345
     REFERENCES	 9-347

10.   TROPOSPHERIC OZONE EFFECTS ON UV-B FLUX AND CLIMATE
     CHANGE PROCESSES	 10-1
     10.1    INTRODUCTION  	 10-1
     10.2    THE ROLE OF TROPOSPHERIC OZONE IN DETERMINING
            GROUND-LEVEL UV-B FLUX	 10-1
            10.2.1     Factors Governing Ultraviolet Radiation Flux at the
                      Earth's Surface 	 10-2
                      10.2.1.1   UV Radiation: Wavelengths, Energies and Depth
                               of Atmospheric Penetration  	 10-2
                      10.2.1.2   Temporal Variations in Solar Flux	 10-3
                      10.2.1.3   Atmospheric Radiative Interactions with Solar
                               UV Radiation  	 10-4
                      10.2.1.4   Modeling Surface UV-B Irradiance	 10-10
            10.2.2     Factors Governing Human Exposure to Ultraviolet Radiation	 10-11
                      10.2.2.1   Outdoor Activities  	 10-12
                      10.2.2.2   Occupation	 10-13
                      10.2.2.3   Age	 10-13
                      10.2.2.4   Gender  	 10-14
                      10.2.2.5   Geography  	 10-15
                      10.2.2.6   Protective Behavior 	 10-15
                      10.2.2.7   Summary of Factors that Affect Human Exposures
                               to Ultraviolet Radiation 	 10-16
            10.2.3     Factors Governing Human Health  Effects due to Ultraviolet
                      Radiation	 10-16
                      10.2.3.1   Erythema	 10-17
                      10.2.3.2   Skin Cancer	 10-19
                      10.2.3.3   Ultraviolet Radiation Exposure and the Incidence
                               of Nonmelanoma Skin Cancers  	 10-20
                      10.2.3.4   Ocular Effects of Ultraviolet Radiation Exposure 	 10-26
                      10.2.3.5   Ultraviolet Radiation and Immune System
                               Suppression	 10-27
                      10.2.3.6   Protective Effects of Ultraviolet Radiation -
                               Production of Vitamin D  	 10-29
            10.2.4     Summary and Conclusions for O3 Effects on UV-B Flux	 10-30
     10.3    TROPOSPHERIC OZONE AND CLIMATE CHANGE 	 10-31
            10.3.1     The Projected Impacts of Global Climate Change 	 10-32
            10.3.2     Solar Energy Transformation and the Components of the
                      Earth's Climate System	 10-36

-------
                                   Table of Contents
                                        (cont'd)
                                                                                   Page
             10.3.3    The Composition of the Atmosphere and the Earth's
                      Radiative Equilibrium	  10-37
                      10.3.3.1    Forcing of the Earth's Radiative Balance	  10-39
             10.3.4    Factors Affecting the Magnitude of Climate Forcing by Ozone	  10-41
                      10.3.4.1    Global versus Regional Atmospheric Ozone
                                Concentrations 	  10-41
                      10.3.4.2    Global Versus Regional Atmospheric Ozone Trends  ....  10-42
                      10.3.4.3    The Sensitivity of Ozone-related Forcing Surface
                                to Albedo 	  10-45
             10.3.5    Estimated Forcing by Tropospheric Ozone	  10-45
                      10.3.5.1    Direct Climate Forcing Due to Ozone  	  10-45
                      10.3.5.2    Indirect Forcing Due to Ozone	  10-47
                      10.3.5.3    Predictions for Future Climate Forcing by
                                Anthropogenic Ozone	  10-48
             10.3.6    Conclusion	  10-48
     REFERENCES	  10-50

11.   EFFECT OF OZONE ON MAN-MADE MATERIALS	  11-1
             11.11.1    Mechanisms of Ozone Damage and Exposure-Response	  11-1
             11.11.2    Textiles andFabrics	  11-3
             11.11.3    Dyes, Pigments, andlnks 	  11-4
             11.11.4    Artists' Pigments	  11-5
             11.11.5    Surface Coatings	  11-8
     REFERENCES	  11-14

-------
                                      List of Tables

Number                                                                             Page

9-1      Advantages and Disadvantages of Protective Chemicals Used in Assessment
         of O3 Effects on Plants  	 9-17

9-2      Advantages and Disadvantages of Bioindicators Used to Study O3 Plant Effects .... 9-18

9-3      Advantages and Disadvantages of Cultivar Comparisons Used in Assessment
         of O3 Effects on Plants  	 9-22

9-4      Advantages and Disadvantages of Various Dendrochronological Techniques
         Used in Assessment of O3 Effects on Plants	 9-24

9-5      Advantages and Disadvantages of Modeling Techniques Used in Assessment
         of O3 Effects on Plants  	 9-27

9-6      The Flow of Ozone into a Leaf and Possible Reactions	 9-39

9-7      Some Rates of Reaction of Ozone With Critical Biochemicals	 9-45

9-8      Superoxide Dismutase Isozymes and Isoforms	 9-62

9-9      Gene Families and cDNA Clones Used as Probes for SAR (Ward et al, 1991) 	 9-66

9-10     Proteins Altered by Ozone as Measured by Molecular Biological Techniques
         as mRNA Level or Other Gene Activity Rather than Enzyme Activity	 9-71

9-11     Interactions Involving O3 and Plant Pathogens	 9-97

9-12     Effects of Increased Carbon Dioxide on Ozone-Induced Responses of Plants
         at the Metabolic, Physiological, and Whole-Plant Levels  	 9-130

9-13     Summary of Ozone Exposure Indices Calculated for 3- or 5-Month Growing
         Seasons from 1982 to 1991	 9-185

9-14     Ozone Exposure Levels (Using Various Indices) Estimated To Cause at Least
         10% Crop Loss in 50 and 75% of Experimental Cases  	 9-187

9-15     SUM06 Levels Associated with 10 and 20% Total Biomass Loss for 50 and
         75% of the Seedling Studies	 9-190

9-16     Summary of Selected Studies of Ozone Effects on Annual  Species   	 9-195

9-17     Summary of Selected Studies of the Effects of Ozone on Perennial
         Herbaceous Plants	 9-199

9-18     Summary of Selected Studies of Ozone Effects on Deciduous Trees and Shrubs .  . . 9-201

-------
                                      List of Tables
                                         (cont'd)

Number                                                                             Page

9-19     Summary of Selected Studies of Ozone Effects on Evergreen Trees and Shrubs  . . .  9-206

9-20     Ethylene Diurea Effects on Vegetation Responses to Ozone	  9-210

9-21     Ozone Exposures at 35 Rural Sites in the Clean Air Status and Trends Network
         in the Central and Eastern United States From 1989 to 1995	  9-228

9-22     Essential Ecological Attributes for Natural Ecosystems Affected by O3	  9-257

9-23     Case Studies Demonstrating the Ecological Effects of O3	  9-264

9-24     The Most Comprehensively Studied Effects of O3 on Natural Ecosystem is
         the San Bernardino Mountain Forest Ecosystem	  9-266

9-25     Effects of Ozone, Ozone and N Deposition, and Ozone and Drought Stress
         on Pinus ponderosa and Pinus jeffreyi in the Sierra Nevada and the San
         Bernardino Mountains, California.  Citations are Focused on Research
         Published Since U.S. EPA (1996)   	  9-267

9-26     Summary of Responses ofPopulits tremuloides to Elevated CO2 (+200 |imol
         mol"1), O3 (1.5 x ambient), or CO2+O3 Compared with Control During Three
         Years of Treatments at the Aspen FACE Project (Modified from Karnosky
         et al. 2003a)  	  9-274

10-1     Examples of Impacts Resulting From Proj ected Changes in Extreme
         Climate Events 	  10-34

10-2     CTM Studies Assessed by the IPCC for its Estimate of the Change in Global,
         Total Column O3 Since the Pre-Industrial Era	  10-44

10-3     Tropospheric O3 Change (HO3) in Dobson Units (DU) Since Pre-industrial
         Times, and the Accompanying net (SW plus LW) Radiative Forcings (Wirf2),
         After Accounting for Stratospheric Temperature Adjustment (using the Fixed
         Dynamical  Heating method)	  10-46

11-1     Average 24-h Ozone Concentrations Producing the Highest Frequency of
         Cracks of a Certain Length in the Middle and Central Zones of the Rubber
         Test Strips	  11-3

11-2     Cuprammonium Fluidity of Moist Cotton Cloth Exposed to 20 to 60 ppb Ozone ....  11-4

11-3     Color Change After 12 Weeks of Exposure to a Mixture of Photochemical
         Oxidants 	  11-11
                                          III-xv

-------
                                      List of Figures

Number                                                                               Page

9-1       Ozone uptake from the atmosphere  	  9-30

9-2       Absorption and transformation of O3 within the leaf	  9-32

9-3       The uptake of O3 into the leaf.  Each of the individual concentration layers of
          O3 represents a different process of movement and of plant/ microenvironmental
          interaction	  9-36

9-4       The microarchitecture of a dicot leaf	  9-37

9-5       The change in the O3 concentration inside a leaf with time	  9-38

9-6       Possible transformations of O3  within a leaf	  9-41

9-7       Possible reactions of O3 within water	  9-41

9-8a,b    The Crigee mechanism of O3 attack of a double bond	  9-44

9-9       Varied ESR radicals, trapped and not, generated by ozone under somewhat
          physiological conditions	  9-47

9-10      Pathogen-Induced Hypersensivity 	  9-50

9-11      The interaction of H2O2 and Ca2+ movements with ABA-induced
          stomatal closure  	  9-52

9-12      The reaction of ascorbate within the apoplasm of the cell wall and its ultimate
          reduction/oxidations	  9-56

9-13      The pathway leading from phospholipids to jasmonic and traumatic acid	  9-70

9-14      The production of Rubisco and its Calvin Cycle pathway reactions 	  9-75

9-15      Linkage of senescence with hypersensitivity reactions and first event of O3
          attack of plants  	  9-82

9-16      Diagrammatic representation of several exposure indices, illustrating how they
          weight concentration and accumulate exposure	  9-162

9-17      Trends in May-September 12-h SUM06, peak 1-h ozone concentration and
          number of daily  exceedances of 95 ppb for Crestline in  1963-1999 in relation
          to trends in mean daily maximum temperature for Crestline and daily reactive
          organic gases (ROG) and oxides of nitrogen (NOX) for San Bernardino county  ....  9-168

-------
                                       List of Figures
                                          (cont'd)

Number                                                                                Page

9-18      Distribution of biomass loss predictions from Weibull and linear
          exposure-response models that relate biomass to O3 exposure	  9-188

9-19      A conceptual diagram of processes and storage pools in sources and sinks that
          are affected by ozone exposure	  9-276

9-20      Common anthropogenic stressors and the essential ecological attributes
          they affect	  9-313

10-1      Extraterrestrial solar flux measured by the satellite UARS SOLSTICE
          instrument (dotted line)	  10-3

10-2      Ozone column abundances from the years 1990 to 1992 for 0, 40, and 80° N
          as well as 80° S	  10-6

10-3      The sensitivity of ground-level UV flux to a 1 DU change in total column
          O3, under clear sky conditions, as a function of solar zenith angle (SZA)	  10-10

10-4      Complexity of factors that determine human exposure to UV radiation  	  10-11

10-5      Estimated global mean radiative forcing exerted by gas and various particle
          phase species for the year 2000, relative to 1750	  10-40

10-6      Mid-tropospheric O3 abundance (ppb) in northern midlatitudes  (36 °N-59 °N)
          for the years  1970 to 1996  	  10-43

11-1      In-service fading of nylon 6 yarn inside house	  11-6

11-2      In-service fading of nylon 6 yarn outside house	  11-7

11-3      Observed color changes for natural colorant-on-paper systems during exposure
          to 0.40 ppm ozone at 25 °C ± 1 °C, 50% RH, in the absence of light  	  11-9

11-4      Observed color changes for natural colorant-on-site during exposure  to
          0.40 ppm ozone at 25 °C ± 1 °C, 50% RH, in the absence of light  	  11-10

-------
                        Authors, Contributors, and Reviewers






To be inserted in Second External Review draft.

-------
                U.S. ENVIRONMENTAL PROTECTION AGENCY
      PROJECT TEAM FOR DEVELOPMENT OF AIR QUALITY CRITERIA
                                FOR OZONE
To be inserted in Second External Review draft.

-------
               U.S. ENVIRONMENTAL PROTECTION AGENCY
              SCIENCE ADVISORY BOARD (SAB) STAFF OFFICE
          CLEAN AIR SCIENTIFIC ADVISORY COMMITTEE (CASAC) OZONE
                                REVIEW PANEL*
To be inserted in Second External Review draft.
                                  III-xx

-------
Abbreviations and Acronyms
AQCD
CxT
CI
CO2
COP
CTM
DNA
DNA
EPA
GSH
H+
H202
HFC
HO
HO2
IPCC
IR
mRNA
NHAPS
NMHC
NO2
NOX
NRC
0('D)
03
OH
PAN
Air Quality Criteria Document

Concentration times duration of exposure
color index
carbon dioxide
Conference of Parties
Chemistry Transport Model
Deoxyribonucleic acid
Deoxyribonucleic acid






U.S. Environmental Protection Agency
Glutathione
hydrogen ion
hydrogen peroxide
hydrofluorocarbon
hydroxyl
hydroperoxyl; hydroperoxy
Intergovernmental Panel on Climate
infrared radiation
Messenger ribonucleic acid






Change


National Human Activity Pattern Survey
nonmethane hydrocarbon
nitrogen dioxide
nitrogen oxides
National Research Council
electronically excited oxygen atom
ozone
hydroxyl; hydroxy
Peroxyacetyl nitrate









-------
RNA              Ribonucleic acid
RO2               organic peroxyl; organic peroxy
SO2                sulfur dioxide
SOD               Superoxide dismutase
SUMOO            Sum of all hourly average concentrations
SUM06            Seasonal sum of all hourly average concentrations > 0.06 ppm
TAR               Third Assessment Report
UNEP             United Nations Environment Program
UV                Ultraviolet
UV-A
UV-B              Ultraviolet radiation of wavelengths from 280 to 320 nm
uv-c
W126              cumulative integrated exposure index with a sigmoidal weighting function
WMO             World Meteorological Organization
ZAPS              Zonal Air Pollution System

-------
 i        9.  ENVIRONMENTAL EFFECTS:  OZONE  EFFECTS
 2                  ON VEGETATION AND ECOSYSTEMS
 3
 4
 5     9.1   INTRODUCTION
 6          The preceding chapters of this document focused on discussion of: (a) background
 7     information regarding ozone (O3)-related atmospheric chemistry, air quality and exposure
 8     aspects; and (b) dosimetric/health effects aspects, as well as the integrative synthesis of key
 9     information drawn from such chapters of most importance for EPA review of the primary O3
10     NAAQS. On the other hand, this and the next two chapters assess available scientific
11     information on O3-related welfare effects of most pertinence for the review of secondary O3
12     NAAQS. This includes discussion of three classes of environmental effects:
13         • O3 effects on vegetation and ecosystem (Chapter 9);
14         • the role of tropospheric O3 in climate change, including determining of ground-level
             UV-B flux (Chapter 10); and
15         • O3 effects on manmade materials (Chapter 11).
16          As for the organizational structure of this chapter, after this opening Introduction,
17     Section 9.2 highlights key features of methodologies used in vegetation research, followed by
18     discussion of vegetation species response/mode-of-action aspects (Section 9.3) and factors that
19     modify functional and growth responses (Section 9.4).  Effects-based air quality exposure and
20     dose-response indices are next discussed in Section 9.5, followed by assessment of O3  exposure-
21     plant response relationships (Section 9.6) and effects on natural ecosystems (Section 9.7). The
22     economic evaluation of O3 effects on agriculture, forestry, and natural ecosystems is then
23     discussed in Section 9.8.  Lastly, the chapter closes with Section 9.9, an overall summary of key
24     findings and conclusions.
25
26
       January 2005                              9-1        DRAFT-DO NOT QUOTE OR CITE

-------
 1      9.2  METHODOLOGIES USED IN VEGETATION RESEARCH
 2      9.2.1  Introduction
 3           The scale of investigations into the direct effects of O3 on plants range from evaluation of
 4      sub-cellular to cellular, organismal, population, community, and ecosystem responses, each with
 5      its particular experimental methodologies and suite of specialized instrumentation, equipment,
 6      facilities, and experimental protocols.  These investigations generate data; whereas other types of
 7      methodologies exist for the handling of data and statistical analysis as well as the utilization of
 8      data in developing the different exposure metrics or indices used to define exposure, quantitative
 9      exposure-response relationships, and computer simulation models of these. The objective of this
10      section is not to provide an updated encyclopedia of all the methods that have been used, but
11      rather to focus on approaches that have:
12            (1)   led to improved understanding of the quantitatively measurable growth and
                   development responses of plants and plant communities to O3, or
13            (2)   provided information about the extent and geographic distribution of the
                   responses of herbaceous and woody plants, both cultivated and native, to
                   ambient O3 exposures.
14      The first objective is essential for determining dose-response functions used in the development
15      of impact and risk assessments of the effects of O3 and usually involves treating plants to a range
16      of artificial O3 exposures.  The second objective is essential for determining the geographic
17      distribution of the risk and usually involves plants subjected to ambient air O3 exposures.
18           The types of methodologies used by biochemists, molecular biologists, or plant
19      physiologists, whose interests lie in determining effects on specific constituents or in
20      understanding the mode of action of O3, are not discussed here. Methods used to characterize
21      the O3 content of ambient air and to define exposure and exposure-response relations are
22      discussed in  Sections 9.5 and 9.6, respectively.
23           The methodologies for exposure-response studies have used many different types of
24      exposure facilities and protocols and have employed a range of statistical approaches to the
25      analysis and interpretation of data. Most of the studies have been conducted using major
26      agricultural crop species. The methodologies have improved over the years as a result of the
27      development, availability, or application of new or improved instrumentation, physical  systems,
28      and numerical approaches to data analysis. Yet equally important to the roles played by these

        January 2005                              9-2        DRAFT-DO NOT QUOTE OR CITE

-------
 1      advances has been the clearer understanding that has emerged from earlier work identifying the
 2      type of experimentation needed to achieve realistic assessments of the magnitude and extent of
 3      the impact of O3 on vegetation  of all types. As a result, significantly increased attention is now
 4      being paid to field observations of responses to ambient O3 pollution, particularly to the
 5      responses of forest trees and native vegetation.
 6          Other than in various exploratory studies that have used chamber-based steady-state
 7      exposure concentrations (so-called "square-wave" exposures), the trend in experimental
 8      exposure protocols has been to attempt to expose plants under conditions as natural as possible
 9      to temporal profiles that simulate the real-world, either by conducting experiments in the field or
10      in elaborately controlled environment facilities that provide simulated field conditions.
11          Previous Air Quality Criteria Documents for Ozone and Other Photochemical Oxidants
12      (U.S. Environmental Protection Agency, 1986; 1996) described the time-course for these
13      methodological developments.  Although this section provides a brief overview of the
14      methodologies used in the past, and their limitations, it focuses mainly on those techniques that
15      have come into prominence over the last decade. This has been aided considerably by several
16      compilations of experimental methodologies and facilities, such as the earlier comprehensive
17      review for the U.S. Environmental Protection Agency/National Acid Precipitation Assessment
18      Program by Hogsett et al. (1987a; 1987b) and the more recent reviews by Manning and Krupa
19      (1992), Musselman and Hale (1997) and Karnosky et al. (2001).
20
21      9.2.2  Methods Involving Experimental Exposures to O3
22      9.2.2A  "Indoor", Controlled Environment, and Greenhouse Chambers
23          The earliest experimental investigations of the effects of O3 on plants utilized simple glass
24      or plastic-covered chambers, often located within greenhouses, into which a flow of O3-enriched
25      air or oxygen could be passed to provide the exposure. The types, shapes, styles, materials of
26      construction, and locations of these chambers were as numerous as the different investigators
27      and, in spite of providing little  resemblance to real-world conditions,  they yielded much of the
28      basic information on the visible and physiological effects on plants. The construction and
29      performance of more elaborate and better instrumented chambers dating back to the 1960s has
30      been well-summarized in Hogsett et al. (1987a), including those installed in greenhouses (with
31      or without some control of temperature and light intensity).

        January 2005                              9-3         DRAFT-DO NOT QUOTE OR CITE

-------
 1           One greenhouse chamber approach that continues to yield useful information on the
 2      relationships of O3 uptake to both physiological and growth effects employs continuous stirred
 3      tank reactors (CSTRs) first described by Heck et al. (1978). Although originally developed to
 4      permit mass-balance studies  of O3 flux to plants, their use has more recently widened to include
 5      short-term physiological and growth studies of O3 x CO2 interactions (e.g., Costa et al., 2001;
 6      Heagle et al., 1994b; Loats and Rebbeck, 1999; Rao et al..  1995; Reinert and Ho. 1995; and
 7      Reinert et al.. 1997), and of surveys of native plant responses to O3 (Orendovici et al., 2003).
 8      In many cases,  supplementary lighting and temperature control of the surrounding structure have
 9      been used to control or modify the environmental conditions, e.g., Heagle et al. (1994a).
10           Many investigations have utilized commercially available controlled environment
11      chambers and walk-in rooms adapted to permit the introduction of a flow of O3 into the
12      controlled air-volume.  Such chambers continue to find use in genetic screening and in
13      physiological and biochemical studies aimed primarily at improving our understanding of mode
14      of action. For example, some of the ongoing studies of the O3 responses ofPlantago major
15      populations have been conducted in controlled environment chambers (Reiling and Davison,
16      1994; Whitfield et al., 1996b).
17           The environmental conditions provided by indoor chambers of any type will always
18      preclude the use of the information obtained with such chambers in predicting O3 effects in the
19      natural environment, because the environmental conditions will always be measurably different
20      from field conditions. However, highly sophisticated controlled environment chambers such as
21      those described by Langebartels et al. (1997), which are subdivided into aerial and root
22      compartments,  with dynamic control of light intensity and photoperiod, air and soil temperature,
23      humidity, soil moisture, wind speed, and exposure to O3, may come close to simulating specific
24      natural conditions.  Such chambers have provided meaningful insights into a wide array of early
25      biochemical responses of plants to O3.  They can minimize confounding factors that make indoor
26      chamber studies only rarely able to be extrapolated to field conditions, i.e., that shoots and roots
27      develop under different temperature regimes.
28           The applicability of the results of many chamber studies is also limited by their use of
29      container-grown plants. Several recent studies have raised serious questions as to the relevance
30      of pot-based studies to the true field situation (even when the studies are conducted in the field.)
31      Most of the questions have concerned studies of the effects of CO2 enrichment, as discussed in

        January 2005                               9-4         DRAFT-DO NOT QUOTE OR CITE

-------
 1      Section 9.4.7.1, but are also relevant to O3 enrichment studies, as shown by Whitfield et al.
 2      (1996a). Whitfield et al., reported significant interactive effects between O3 and soil volume on
 3      the growth ofPlantago major. They noted that although container size may limit root and,
 4      hence, plant growth, the reverse may also be true for single plants in large containers which do
 5      not experience typical field competition for resources.
 6
 7      9.2.2.2 Field Chambers
 8           Although closed field chambers have largely fallen out of favor in recent years, closed
 9      "Solardome" field chambers (Lucas et al., 1987; Rafarel and Ashenden, 1991) have been used
10      recently in studies of O3 x acid mist interactions (Ashenden et al., 1995; 1996).
11           Concern over the need to establish realistic plant-litter-soil relationships as a prerequisite to
12      studies of the effects of O3 and CO2 enrichment on ponderosa pine (Pinus ponder osa) seedlings
13      led Tingey et al. (1996) to develop  closed, partially environmentally controlled, sun-lit chambers
14      ("terracosms") incorporating 1 m-deep lysimeters containing forest soil that retained the
15      appropriate horizon structure.
16           In general, field chamber studies are dominated by the use of various versions of the open-
17      top chamber (OTC) design, first described by Heagle et al. (1973) and  Mandl et al. in 1973.
18      Most chambers are ~3 m in diameter with 2.5 m high walls. Hogsett et al. (1987a) described in
19      detail many of the various modifications to the original OTC designs that appeared subsequently,
20      e.g., the use of larger chambers to permit exposing small trees (Kats et al., 1985); and grapevines
21      (Mandl et al.,  1989), the addition of a conical baffle at the top to improve ventilation (Kats et al.,
22      1976), a frustrum at the top to reduce ambient air incursions, and a plastic rain-cap to exclude
23      precipitation (Hogsett et al., 1985). All of these modifications included the discharge of air via
24      ports in annular ducting or interiorly perforated double-layered walls at the base of the chambers
25      to provide turbulent mixing and upward mass flow of air.
26           Wiltshire et al. (1992) described a large open-top chamber suitable for small trees with
27      roll-up sides that permitted the trees to be readily subjected from time to time to episodic,
28      normal, "chamberless" environmental  conditions. In the 6 m-high OTCs described by Seufert
29      and Arndt (1985) used with Norway spruce (Picea abies) trees, a second zone of annular
30      enrichment was also provided between 4 and 5 m.  The use of OTCs was adopted for the large
31      European Stress Physiology and Climate Experiment on the effects of CO2 and O3 on spring

        January 2005                               9-5        DRAFT-DO NOT QUOTE OR CITE

-------
 1      wheat (ESPACE-wheat), conducted in 1994-1996 at field sites in eight countries (Jager et al.,
 2      1999). However, typical European chambers have the introduction of O3-enriched air at or
 3      above canopy height. The relatively low costs of fabrication, operation and maintenance has
 4      favored OTC use in field studies (Fangmeier et al., 1992; Musselman and Hale, 1997).  The air
 5      supplied to the chambers can be readily filtered through activated charcoal to reduce the O3
 6      concentration or it can be enriched with O3 to provide a range of exposures.
 7           All field chambers create internal environments that differ from ambient air, giving rise to
 8      so-called "chamber effects" with the modification of microclimatic variables, including reduced
 9      and uneven light intensity, uneven rainfall, constant wind speed, reduced dew formation, and
10      increased air temperatures (Fuhrer, 1994; Manning and Krupa, 1992).  Several shortcomings of
11      the OTC design and operation relate to the means of introduction and mixing of enriched air to
12      produce a definable exposure.  First, the plants are subjected to constant turbulence, which,
13      through increased uptake resulting from the consequently low boundary layer resistance to
14      diffusion, may lead to overestimation of the magnitude of real-world cause-effect relationships
15      (Krupa et al., 1995; Legge et al., 1995). Second, the introduction of the O3-enriched air in the
16      lower part of chambers described by Heagle et al. (1973) and Mandl et al. (1973) results in a O3
17      concentration gradient that decreases with increasing height, the converse of the situation
18      observed in ambient  air, in which the O3 concentration decreases markedly from above a plant
19      canopy to ground level (Griinhage and Jager, 1994; Pleijel et al., 1995; Pleijel et al., 1996).
20      Concern that studies  conducted in such OTCs may somewhat overestimate the effects of O3 led
21      to the European design which provides a decreasing downward gradient. It seems unlikely that
22      the "chamber effects" produced by the two designs will be the same.  These issues are discussed
23      more fully in Section 9.2.2.4.
24           It should also be noted that, although originally developed for exposing row drops in the
25      field, many recent studies employing OTCs have used potted plants in order to include or control
26      edaphic or nutritional factors or water relations within the experimental design. Therefore, the
27      same caveats as those discussed above (Section 9.2.2.1) with regard to extrapolating results of
28      pot studies to true field conditions apply to OTC studies.
29           The difficulties faced in the experimental  exposure of forest trees to air pollutants in
30      chambers (e.g., Seufert and Arndt [1985]) led to the development of branch chambers such as
31      those described by Ennis et al. (1990), Houpis et al. (1991) and Teskey et al. (1991).  These

        January 2005                               9-6        DRAFT-DO NOT QUOTE OR CITE

-------
 1      chambers are essentially large cuvettes and, as noted by Musselman and Hale (1997), share
 2      many of the characteristics of CSTRs, i.e., transparent walls, internal fans, and inlet and outlet
 3      monitors to permit the determination of O3 uptake, CO2 exchange, and transpiration. Although
 4      they make it  possible to expose whole branches to different O3 regimes, the relevance of the data
 5      they yield in  regard to the whole tree may be questioned. As noted by Saxe et al. (1998), the
 6      inevitable change in environmental conditions resulting from the isolation of the branch may
 7      cause different responses from those that would be obtained if the whole tree was subjected to
 8      the same environmental conditions.
 9
10      9.2.2.3 Plume Systems
11           Plume systems are chamberless exposure facilities in which the atmosphere surrounding
12      plants in the field is modified by the injection of pollutant gas into the air above or around them
13      from multiple orifices that are spaced to  permit diffusion and turbulence so as to establish
14      relatively homogeneous conditions as the individual plumes disperse and  mix with the ambient
15      air.  As pointed out by Manning and Krupa (1992), they can only be used to increase the O3
16      levels in the ambient air. The volume of air to be modified is unconfined, and three approaches
17      have been used to achieve desired pollutant concentrations in the air passing over the plants
18      producing various systems that:
19           (1)  achieve a concentration gradient, in most instances dependent upon the direction of
                  the prevailing wind;
20           (2)  achieve spatially uniform concentrations over a plot, dependent upon wind
                  direction; and
21           (3)  seek to achieve spatially uniform concentrations over a plot, independent of wind
                  speed and direction.
22           Gradient systems created by dispensing a pollutant gas into the air at canopy level from
23      perforated horizontal pipes arranged at right angles to the prevailing wind were described for
24      SO2 studies in the early  1980s. A modified gradient system for O3 was used by Bytnerowicz
25      et al. (1988) to study effects on desert species, but there appear to have been no recent
26      applications of the method.  A gradient O3-exclusion system is discussed in Section 9.2.3.1.
27           Systems designed to achieve spatially uniform pollutant levels by ensuring that the release
28      of a pollutant is always on the upwind side of the study site were also originally described for

        January 2005                              9-7         DRAFT-DO NOT QUOTE OR CITE

-------
 1      SO2 studies, e.g. Greenwood et al. (1982). However, the adaptation of these concepts as
 2      introduced by McLeod et al. (1985) in constructing a large circular field site for exposing crops
 3      to SO2 led to the subsequent development of both the large-scale O3 and SO2 fumigation system
 4      for forest trees in the United Kingdom in 1985 (the Liphook Forest Fumigation Project; McLeod
 5      et al., 1992), the smaller system for O3 fumigation constructed at Kuopio, Finland in 1990
 6      (Wulff et al., 1992), and the free-air carbon-dioxide enrichment (FACE) systems of gas dispersal
 7      over crops (Hendrey and Kimball, 1994) and forest trees (Hendrey et al., 1999).  Although
 8      originally designed to provide chamberless field facilities for studying the CO2 effects of climate
 9      change, large forest tree FACE systems have recently been adapted to include the dispensing of
10      O3 (Karnosky et al., 1999).  Volk et al. (2003) have recently described a system for exposing
11      grasslands that uses 7-m diameter plots. FACE systems discharge the pollutant gas (and/or CO2)
12      through spaced orifices along an annular ring (or torus) or at different heights on a ring of
13      vertical pipes. Computer-controlled feedback from the monitoring of gas concentration
14      regulates the feed rate of enriched air to the dispersion pipes. Feedback of wind speed and
15      direction information ensures that the discharges only occur upwind of the treatment plots, and
16      that discharge is restricted or closed down during periods of low wind speed or calm conditions.
17      The diameter of the arrays and their heights in some FACE systems (25-30 m) requires large
18      throughputs of enriched air per plot, particularly in forest tree systems. The cost of the
19      throughputs tends to limit the number of enrichment treatments, although Hendrey et al.  (1999)
20      have argued that the cost on an enriched volume basis is comparable to that of chamber systems.
21          An alternative to the FACE system to free-air fumigation uses a horizontal grid system
22      through which pollutant-air enriched is discharged over the canopies of plants in field plots.  The
23      original design, termed the Zonal Air Pollution System (ZAPS), was developed for studying the
24      effects of SO2 on native grasslands (Lee et al.,  1975), and it was later modified by Runeckles
25      et al. (1990) by randomly dividing each of three treatment plots into four sub-plots, each with
26      different numbers of discharge orifices to provide various levels of O3 enrichment.  With the
27      ZAPS system, changes in wind direction and speed result in varying  degrees of carry-over from
28      sub-plot to sub-plot, effectively resulting in twelve stochastically different seasonal exposures.
29      The system was used for studies of growth effects on field crops and 2- to 4-year old Douglas fir
30      (Pseudotsuga menziesii) saplings (Runeckles and Wright, 1996).
        January 2005                               9-8        DRAFT-DO NOT QUOTE OR CITE

-------
 1           A larger ZAPS design was used by Wilbourn et al. (1995) on a grass (Loliumperenne)-
 2      clover (Trifolium repens) mixture and by Ollerenshaw et al. (1999) on oilseed rape (Brassica
 3      napus), whereby four replicate field plots were exposed to intermittent constant additions of O3
 4      to ambient air.  A ZAPS design with eight spatially separated treatment plots was also developed
 5      to obtain crop response data used in assessing crop losses in the Fraser Valley, British Columbia,
 6      Canada (Runeckles and Bowen, 2000).
 7           The FACE-type facility developed for the Kranzberg Ozone Fumigation Experiment in
 8      Germany that began in 2000 (KROFEX; Werner and Fabian, 2002; Nunn et al., 2002) to study
 9      the effects of O3 on mature stands of beech (Fagus sylvaticd) and spruce  (Picea abies) trees is
10      more truly a zonal  system that functions independently of wind direction. The enrichment of a
11      large volume of the ambient air immediately above the canopy takes place via orifices in vertical
12      tubes suspended from a horizontal grid supported above the canopy.
13           Recognizing  the difficulties of modifying the aerial environments of large trees, Tjoelker
14      et al. (1994) devised a free-air system for exposing branches of sugar maple (Acer saccharum)
15      trees to O3. Near the ends of up to 10 branches, enriched air was discharged through small holes
16      in 38-cm-diameter loops of 0.635-cm-OD teflon tubes positioned 20-30 cm below the terminal
17      foliage cluster.
18           Although plume systems make virtually none of the modifications to the physical
19      environment that are inevitable with chambers, their successful use depends on selecting the
20      appropriate numbers, sizes, and orientations of the discharge orifices to avoid "hot-spots"
21      resulting from the direct impingement of jets of pollutant-enriched air on plant foliage (Werner
22      and Fabian, 2002). However, because mixing is unassisted and completely dependent on wind
23      turbulence and  diffusion, local gradients are inevitable even in large-scale FACE systems.  Both
24      FACE and ZAPS systems have provisions for shutting down under low wind speed or calm
25      conditions and for  an experimental area that is usually defined within a generous border, in order
26      to strive for homogeneity of the exposure concentrations within the treatment area. They are
27      also both dependent upon continuous computer-controlled feedback of the O3 concentrations in
28      the mixed treated air and the meteorological conditions.
29
30
        January 2005                              9-9         DRAFT-DO NOT QUOTE OR CITE

-------
 1      9.2.2.4 Comparative Studies
 2           All experimental approaches to the exposure of plants to O3 have shortcomings. The use of
 3      laboratory, greenhouse, or field chambers raises concerns for the roles of "chamber effects" on
 4      micrometeorology as well as the consistent turbulence over and within the plant canopy during
 5      chamber operation, in modifying plant responses.
 6           With the widespread use of the OTC, numerous studies have demonstrated increased air
 7      temperatures and decreased light intensities as regularly noted features of their use. However,
 8      the question to be answered is whether or not these differences affect plant response to O3.
 9      As noted in the 1996 criteria document (U.S. Environmental Protection Agency, 1996), evidence
10      from the comparative studies of OTCs and from closed chamber and O3-exclusion exposure
11      systems on the growth of alfalfa (Medicago saliva) by Olszyk et al.  (1986a) suggested that, since
12      significant differences were found for fewer than 10% of the growth parameters measured, the
13      responses were, in general, essentially the same regardless  of exposure system used and
14      "chamber effects" did not significantly affect response. In  1988, Heagle et al. (1988) concluded:
15      "Although chamber effects on yield  are common, there are no results showing that this will
16      result in a changed yield response to O3." Several more recent studies have, however, indicated
17      that the temperature effect alone may be sufficient to cause significant (though modest) shifts in
18      quantitative responses to O3, both positive and negative, as discussed below in Section 9.4.3.2.
19      Still, it is not clear whether these effects are directly related to temperature or are the result of
20      temperature interactions with other environmental variables.  For example, Olszyk et al. (1992)
21      undertook a 3-year study  of the impact of O3 on Valencia orange trees in large OTCs to
22      determine if "insidious differences in microclimatic conditions could alter plant growth
23      responses and susceptibility to pollutant stress."  Non-filtered chambers were found to have
24      somewhat lower average O3 concentrations than the ambient air, with fewer hourly exceedances
25      of 100 ppb. In cool seasons, stomatal conductance was also lower, implying lower O3 uptake.
26      However, the cumulative fruit yields were doubled in the chamber trees even though
27      photosynthetically active radiation was consistently reduced by about 19%, while leaf
28      temperatures averaged more than 2°C higher.  These data may be somewhat extreme, but they
29      emphasize the need to avoid automatically assuming that OTCs yield response data that are all
30      immediately relevant to the real world, particularly since, as in this study, no O3 enrichment was
31      involved as a complicating factor.

        January 2005                              9-10        DRAFT-DO NOT QUOTE OR CITE

-------
 1           Plume systems avoid "chamber effects," but because they rely solely upon diffusion and
 2      natural turbulence to modify the ambient O3 concentration, they may fail to achieve homogeneity
 3      of the air to which the plants are exposed and may give rise to "hot spots" in which the enriched
 4      air jets are inadequately diluted and impinge directly on foliage. A further deterrent to their
 5      widespread use is the large-scale generation of O3 needed, which has in most cases limited the
 6      numbers of treatments that can be included in an experimental design.  In spite of the various
 7      advantages and disadvantages of the two field systems, it appears that no direct comparative
 8      studies have been conducted.
 9
10      9.2.2.5 Ozone Generation Systems
11           Two approaches have been used to generate the O3 needed for enrichment purposes from
12      air or oxygen: (1) high voltage static discharge and (2) high intensity UV-irradiation.  Using
13      gaseous oxygen as feedstock, both generate O3-enriched  oxygen, free from other impurities.
14      However, the use of high voltage discharge equipment with air  as feedstock requires that the
15      output be scrubbed with water to remove appreciable amounts of the higher oxides of nitrogen
16      (especially N2O5, nitric acid vapor) that form concurrently with O3 (Brown and Roberts, 1988;
17      Taylor et al., 1993).
18
19      9.2.2.6 Experimental Exposure Protocols
20           A few recent chamber studies of physiological or biochemical effects have continued to
21      use "square-wave" exposure profiles typified by a rapid rise to and falling off from a steady
22      target concentration. However, most recent experimentations into O3 effects  on plant growth and
23      development have employed either simulations  of the diurnal ambient O3  profile or
24      enhancement/reduction of the ambient O3 concentrations.
25           Although Hogsett et al. (1985), Lefohn et  al. (1986), and others have described the use in
26      controlled chambers of daily exposure profiles based on observed ambient O3 profiles and such
27      profiles were used in the elaborately controlled  chamber studies of Langbartels et al. (1997),
28      several recent chamber studies have used simpler computer-controlled half- or full-cosine wave
29      profiles to simulate the typical  daily rise and fall in ambient O3  levels (McKee et al., 1997a;
30      1997b; Mazarura, 1997).
        January 2005                              9-11        DRAFT-DO NOT QUOTE OR CITE

-------
 1           The early studies with OTCs involved adding constant levels of O3 to ambient air O3
 2      concentrations, but all recent studies have used enrichment delivery systems that maintain
 3      proportionality to and track ambient O3 concentrations to produce levels that more closely
 4      resemble field observations.  Both FACE and ZAPS studies have used proportional enrichment
 5      to provide a range of treatments, although Wilbourn et al. (1995) and Ollerenshaw et al. (1999)
 6      adjusted their systems manually to obtain a relatively constant target concentration during
 7      exposure episodes.
 8
 9      9.2.3  Methods Involving Exposures to O3 in Ambient Air
10           The experimental methods discussed above are largely aimed at developing quantitative
11      growth-response functions to permit the estimation of the effects of different ambient O3
12      scenarios. Because such methodologies usually involve exposures to higher than ambient O3
13      levels, the applicability of the functions obtained may, to some extent, be relevant only to
14      locations that are naturally subjected to high ambient O3 levels.  Furthermore, as pointed out in
15      Section 9.4, the response functions that they generate rarely incorporate other environmental,
16      genetic, and physiological factors, many of which can severely modify the magnitude of the
17      response to O3. The consequences of ignoring such modifications have been well stated by De
18      Santis (1999).  The European standard for protecting crops (based on the AOT40 index) was
19      derived from OTC studies of O3-induced grain loss of wheat observed  in experiments conducted
20      mostly in non-Mediterranean locations.  However, the impact of ambient O3 on wheat yields in
21      the Po Valley of northern Italy is much less than the devastatingly high loss (> 60%) suggested
22      by the seasonal exceedances of the standard.  On a similar note, Manning (2003) has recently
23      urged the absolute necessity of seeking "ground truth" as verification of the nature and
24      magnitude of impacts on vegetation suggested by response functions using ambient O3
25      monitoring data.
26           Such concerns clearly show that attention needs to be focused on incorporating
27      consideration of environmental and other factors into the response functions upon which
28      standards are based.  This will require the development of improved simulation response models.
29      These concerns have also led to increasing attention being paid to seeking and developing
30      alternative approaches to the assessment of impact and the geographic extent of such impact,
31      approaches that are based on in situ exposures to ambient or sub-ambient O3 levels.  Although

        January 2005                              9-12        DRAFT-DO NOT QUOTE OR CITE

-------
 1      one approach, the use of air exclusion systems, requires experimental facilities, the other
 2      approaches are generally based on simple field observations or measurements and, hence, can be
 3      undertaken on a wide geographic scale.
 4
 5      9.2.3.1 Air-exclusion Systems
 6           The term, air-exclusion system, usually refers to chamberless field systems specifically
 7      designed to protect plants from exposure to polluted air by blowing filtered air through their
 8      canopies. Hogsett et al. (1987a;  1987b) describe several dedicated systems developed in the
 9      1960s and 1970s, but there appear to have been no recent O3-exclusion studies using systems
10      specifically designed for the purpose since those described by Olszyk et al. (1986a; 1986b).
11      Their system, a modification of the earlier system of Jones et al. (1977), consisted of perforated
12      31.8-cm OD  inflatable polyethylene tubes laid between crop rows and supplied with charcoal -
13      filtered air. By increasing the size of the orifices progressively in sections along the 9-m length
14      of the tubes,  an exclusion gradient was  created with a progressive decrease in O3 levels in the air
15      surrounding the crop from one end of the system to the other. The system was used for studies
16      on alfalfa (Medicago sativaL.) comparing plant response in OTCs, closed field chambers, the
17      air-exclusion system, and ambient air plots (as discussed above in Section 9.2.2.4).
18           An air-exclusion component has also been part of the overall design of the many OTC
19      experiments. Charcoal-filtered air or mixtures of charcoal-filtered and ambient air to chambers
20      as part of the overall design.
21
22      9.2.3.2 Natural Gradients
23           Naturally occurring locational differences in ambient O3 concentrations hold potential for
24      the examination of plant response along a gradient of such concentrations. However, few such
25      gradients can be found which meet the rigorous statistical requirements for comparable site
26      characteristics such as soil type, temperature, rainfall, radiation, and aspect (Manning and Krupa,
27      1992), although with small plants, soil variability can be avoided by the use of potted plants.
28           Studies in the 1970s used the natural gradients occurring in Southern California to assess
29      yield losses of alfalfa (saliva) and tomato (Lycopersicon esculentum L.} (Oshima et al., 1976;
30      1977).  A transect study of the impact of O3 on the growth of clover (Trifolium repens L.} and
31      barley (Hordeum vulgar e L.) in the United Kingdom was confounded by differences in the

        January 2005                              9-13        DRAFT-DO NOT QUOTE OR CITE

-------
 1      concurrent gradients of SO2 and NO2 pollution (Ashmore et al., 1988). Studies of forest tree
 2      species in national parks in the eastern U.S. (Winner et al., 1989) revealed increasing gradients
 3      of O3 and visible foliar injury with increased elevation.
 4
 5      9.2.3.3 Use of Chemical Protectants
 6           The use of protective chemicals is a relatively inexpensive, promising alternative to
 7      experimental field exposures in chambers or free-air systems for determining plant response to
 8      O3.  Several chemical compounds (antioxidants, antisenescence agents, fungicides, pesticides,
 9      etc.) have been known for many years to provide plants some protection from photochemical
10      oxidants, such as O3 (Manning and Krupa, 1992). Most of these chemicals were originally used
11      as a one-time application to reduce visible injury caused by acute O3 exposures.  The most
12      widely used and popular of these has been ethylenediurea (EDU). Carnahan et al. (1978)
13      reported that EDU protected pinto bean (Phaseolus vulgaris) from acute O3 injury. After this
14      initial investigation, EDU was shown to suppress visible O3 injury on several species of plants
15      under both controlled and field conditions (Brennan et al., 1987; Clarke et al.,  1983). However,
16      due to lack of a commercial market for this product, its commercial manufacture was largely
17      discontinued. Other chemicals, including benomyl (Manning et al., 1974), carboxin (Rich et al.,
18      1974), ascorbic acid (Dass and Weaver, 1968) and others (Manning and Krupa, 1992), exhibited
19      some beneficial  effects in reducing visible O3 injury.
20           Despite their widespread use for O3-effects studies,  OTCs are not effective in determining
21      plant responses in the field under truly natural conditions, as discussed above,  and are not
22      efficient or easily used in unmanaged ecosystems such as wilderness areas (Manning and Krupa,
23      1992). For these reasons, a renewed interest in using protective chemicals as research tools
24      occurred in the early 1990s.  These chemicals (mostly EDU) have recently been used in studies
25      of different plant species, both in the United States (Bergweiler and Manning,  1999; Kuehler and
26      Flagler, 1999) and in Europe (Bortier et al., 2001b; Pleijel et al., 1999; Wu and Tiedemann,
27      2002), to determine if ambient O3 concentrations affect plant growth and productivity or are just
28      exacerbating foliar injury. For example, Bortier et al.  (2001a) injected seedlings of an O3-
29      sensitive poplar  (Populus nigra) clone with EDU and measured growth over a  1-year period at a
30      field site near Brussels, Belgium.  Over the growing season, stem diameter increment was
31      significantly higher (16%), biomass  was increased (9%), and foliar O3 symptoms were slightly

        January 2005                             9-14        DRAFT-DO NOT  QUOTE OR CITE

-------
 1      less for the EDU-treated seedlings. Ozone levels were reported to be low (AOT40 = 6170 ppb.h,
 2      May-September) during the exposure period. In another study, Manning et al. (2003) applied
 3      EDU (foliar spray) and sodium erythorbate (NaE) at various concentrations, biweekly for three
 4      growing seasons to loblolly pine (Pinus taedd) at a field site in east Texas.  After 3 years, the
 5      trees were harvested and biomass measured. Neither EDU nor NaE prevented foliar O3 injury,
 6      but EDU applications at 450 ppm resulted in increases in stem diameter and height and total
 7      above-ground biomass.  These measures of growth also tended to slightly increase with
 8      applications of NaE, but the effects were nonsignificant.
 9           Several recent studies used EDU in assessing the response of several plant species to O3 to
10      help validate the proposed critical level (AOT40 = 3000 ppb.h) for crop protection in Europe
11      (Ball et al., 1998; Ribas and Penuelas, 2000; Tonneijck and Van Dijk, 2002b; 2002a).  EDU
12      appeared to provide protection from visible foliar injury, but the results regarding yield and
13      biomass reductions were mixed. In a 3-year study over 12 sites throughout Europe, Ball et al.
14      (1998)  used the ratio of EDU-treated versus non-treated white clover (Trifolium repens)
15      biomass but did not find a significant relationship between biomass reductions and AOT40.
16      When other parameters, such as temperature and VPD were included in the model however,
17      they found a significant relationship (r2 = 0.79) between biomass reductions and AOT40, and
18      the greatest biomass reductions occurred in areas with the highest levels of industrialization
19      (in Germany).
20           Tonneijck and Van Dijk (2002b) assessed the relationship of visible injury of subterranean
21      clover (Trifolium subterraneum) to ambient O3 at four sites over three growing seasons in the
22      Netherlands, using EDU-treated and non-treated plants. Visible injury varied by site and year,
23      but was reduced to near zero by EDU treatment. However, no relationship indicative of a
24      protective effect of EDU with this plant species was observed for biomass.  Tonneijck  and
25      Van Dijk (2002a) also reported similar results with bean (Phaseolus vulgaris). Both EDU-
26      treated and non-treated plants were exposed to ambient O3 at three locations in Spain over one
27      growing season (Ribas and Penuelas,  2000). Reductions in yield and biomass were correlated
28      with O3 concentration and EDU provided some protective effect, although results varied by
29      location and with meteorological conditions.
30           The mechanisms by which protective chemicals, especially EDU, protect plants are poorly
31      understood. However, Wu and von Tiedemann (2002) reported that applications of two recently

        January 2005                              9-15        DRAFT-DO NOT QUOTE OR CITE

-------
 1      developed fungicides (azoxystrobin and epoxiconazole) provided protection of spring barley to
 2      relatively high O3 exposures (150-250 ppb, 5 days, 7h/d) and resulted in increases in leaf soluble
 3      protein content, as well as the activity of several antioxidative enzymes (e.g., superoxide
 4      dismutase, catalase, ascorbate-peroxidase, and glutathione reductase). In addition, the increase
 5      in these enzymes reduced superoxide levels in the leaves.
 6           Before using these chemicals in a field setting, preliminary investigations should be done
 7      to evaluate the methods and timing of application, as well as proper application rates, so as to
 8      avoid any potential toxic effects (Manning, 2000; Manning and Krupa, 1992). Also, because
 9      certain plant species may be nonresponsive, careful pre-screening of various species to be tested
10      is critical.
11           In summary, the use of protectants such as EDU (see Table 9-1) provides a ready means of
12      demonstrating the local occurrence of adverse effects of O3, both in terms of visible injury and
13      growth reduction. However,  even if accompanied by some type of O3 and meteorological
14      monitoring, their use has yet to give rise to a methodology for quantitatively assessing exposure
15      response.
16
17      9.2.3.4 Biomonitoring
18      Bioindicators
19           The use of biological indicators to detect the presence of O3 injury to plants is a
20      longstanding and effective methodology (Chappelka and Samuelson, 1998; Manning and Krupa,
21      1992).  A bioindicator can be defined as a vascular or non-vascular plant exhibiting a typical and
22      verifiable response when exposed to a plant stress such as an air pollutant (Manning et al., 2003).
23      To be considered a good indicator species, plants must:
24            (1)   exhibit a distinct, verified response;
25            (2)   have few or no confounding disease or pest problems; and
26            (3)   exhibit genetic stability.
27           Such sensitive plants can be used to detect the presence of a specific air pollutant such as
28      O3 in the ambient air at a specific location or region and, as a result of the magnitude of their
29      response, provide unique information regarding specific ambient air quality. Bioindicators can
30      be either introduced sentinels, such as the widely used tobacco (Nicotiana tabacum) variety

        January 2005                              9-16        DRAFT-DO NOT QUOTE OR CITE

-------
          Table 9-1.  Advantages and Disadvantages of Protective Chemicals Used in Assessment
                                         of O3 Effects on Plants
         Advantages
         No chambers required. Plants exposed to ambient conditions of O3, light, temperature, etc.
         Can conduct studies "in situ."  Equipment needs are minimal. No "chamber effects"
         A high degree of replication possible both within and among locations
         Disadvantages
         Exposure-response studies require inclusion of other methodologies (OTCs, etc.)
         Need measurements of ambient O3 and other meteorological variables (temp, rainfall, etc)
         Have to conduct preliminary toxicology studies to determine proper rate, timing etc.
         Possible plant toxicity can result from repeated applications
         Species response can vary; need to screen for proper species to use
         Modified from Manning and Krupa (1992).
 1     Bel W3, or detectors, which are sensitive native plant species. The approach is especially useful
 2     in areas where O3 monitors are not operated (Manning et al., 2003). For example, in remote
 3     wilderness areas where instrument monitoring is generally not available, the use of bioindicator
 4     surveys in conjunction with the use of passive samplers (Krupa et al., 2001) is a particularly
 5     useful methodology (Manning et al., 2003). However, the method requires expertise or training
 6     in recognizing those signs and symptoms uniquely attributable to exposure to O3 as well as their
 7     quantitative assessment.
 8           Since the 1996 O3 AQCD (U.S. Environmental Protection Agency, 1996) many new
 9     sensitive species have been identified from controlled exposure studies and verified in the field
10     (Flagler, 1998; Innes  et al., 2001). In addition, several new uses of this methodology have been
11     demonstrated, including a national O3 bioindicator network, studies in wilderness areas, and
12     mature tree studies. Although it has been difficult to find robust relationships between the foliar
13     injury symptoms caused by O3 and effects on plant productivity or ecosystem function, visible
14     injury correlations with growth responses have been reported (Table 9-2; Chappelka and
15     Samuelson, 1998; Manning et al., 2003; Smith et al., 2003). One workshop on the utility of
16     bioindicators of air pollutants led to a useful series of peer-reviewed publications in
17     Environmental Pollution (Skelly, 2003).

       January 2005                               9-17        DRAFT-DO NOT QUOTE OR CITE

-------
                Table 9-2. Advantages and Disadvantages of Bioindicators Used to Study
                                            O3 Plant Effects
         Advantages
         No chambers required.  Plants exposed to ambient conditions of O3, light, temperature, etc.
         Relatively inexpensive. Equipment needs are minimal. No "chamber effects"
         A high degree of replication possible (sentinels) both within and among locations
         Disadvantages
         Individuals need to be trained and experienced in O3 symptom recognition
         Need adequate numbers of plants (detectors) to ensure valid results
         Need preliminary tests to insure a constant symptomotology of material used
         Use more than one indicator species (detector) per area if possible
         Quantify site characteristics (soils, light) that may influence symptom expression
         Need measurements of ambient O3 (active or passive) and other meteorological variables
         (temp, rainfall, etc)
         Need to ensure cultural (sentinels) practices (soil, irrigation, fertilization, etc.) are similar
         among sites
 1           National network
 2           The U.S. Forest Service in cooperation with other federal and state agencies developed a
 3      network of O3 bioindicators to detect the presence of O3 in forested systems throughout the U.S.
 4      (Smith et al., 2003). This ongoing program was initiated in 1994; and 33 states currently
 5      participate. In a coordinated effort, a systematic grid system is used as the basis of plot
 6      selection, and field crews are trained to evaluate O3 symptoms on sensitive plant species within
 7      the plots (Coulston et al., 2003; Smith et al., 2003).
 8           The network has provided evidence of O3 concentrations high enough to induce visible
 9      symptoms on sensitive vegetation. From repeated observations and measurements made over a
10      number of years, specific patterns of areas experiencing visible O3 injury symptoms can be
11      identified. Coulston et al. (2003) used information gathered over a 6-year period (1994-1999)
12      from the network to identify several species that were sensitive to O3 over a regional scale
        January 2005                              9-18        DRAFT-DO NOT QUOTE OR CITE

-------
 1      included sweetgum (Liquidambar styraciflua), loblolly pine (Pinus Taeda), and black cherry
 2      (Prunus serotind).
 3
 4           Wilderness areas
 5           The use of bioindicator species as detectors has proven to be an effective technique for
 6      deriving a relative estimate of O3 injury in wilderness areas in both the US and Europe
 7      (Chappelka et al., 1997; 2003; Manning et al., 2002). However, to be truly effective, these
 8      regional and national bioindicator studies need the inclusion of air quality data and related
 9      growth studies to determine effects on productivity and ecosystem function (Bytnerowicz et al.,
10      2002; Manning et al., 2003; Smith et al., 2003).
11           Chappelka et al. (1997; 2003) conducted surveys of foliar injury on several native plant
12      species throughout the Great Smoky Mountains National Park (GRSM), including black cherry
13      (Prunus serotind), tall milkweed (Ascelpias exaltatd), cutleaf coneflower (Rudbeckia laciniatd)
14      and crownbeard (Verbesina occidentalis). Visible foliar symptoms were prevalent throughout
15      the Park, indicating that injury-producing O3 levels were widespread in GRSM.
16           Manning et al. (2002) recently summarized a multi-year (1993-2000) bioindicator project
17      in the Carpathian Mountain range in eastern Europe. They evaluated numerous trees, shrubs,
18      forbs, and vines for possible symptoms of O3 injury.  Observations were made at plots located in
19      the vicinity of either active or passive O3 monitors (Bytnerowicz et al., 2002). Approximately 30
20      species of native plants detectors were identified as possible bioindicators, the majority of which
21      (21) were shrubs (Manning et al., 2002). Based on these observations, it was concluded that O3
22      concentrations were sufficiently high to impact ecosystems in the region. Similar investigations
23      regarding the sensitivity of native species have been conducted in  Switzerland (Novak et al.,
24      2003) and Spain (Orendovici et al., 2003).
25
26           Mature tree detectors
27           Many studies have reported visible injury of mature coniferous trees caused by O3,
28      primarily in the western United States  (Arbaugh et al., 1998) and, to a lesser extent, to mature
29      deciduous trees in eastern North America. In an effort to determine the extent and magnitude of
30      visible injury in mature tree canopies, Hildebrand et al. (1996) and Chappelka et al. (1999b)
31      conducted independent studies in the GRSM and the Shenandoah National Park (SHEN). The

        January 2005                             9-19        DRAFT-DO NOT QUOTE OR CITE

-------
 1      species examined were sassafras (Sassafras albiduni), black cherry, and yellow-poplar
 2      (Liriodendron tulipifera L.) in GRSM and white ash (Fraxinus americanaZ.), black cherry, and
 3      yellow-poplar in SHEN.  Protocols were similar at both parks, and trees were located near O3
 4      monitors at three different areas in each park. Results from both studies indicated that symptoms
 5      of O3 injury were present in the trees and correlated with the amount of injury both spatially and
 6      temporally. Ozone injury tended to be most severe at the highest elevation, except with yellow-
 7      poplar.
 8           Hildebrand et al. (1996) observed significant O3 exposure-plant response relationships with
 9      black cherry. The best relationships were found between foliar injury and the SUM06 and W126
10      exposure indices, indicating that higher O3  concentrations were important in eliciting a response
11      in black cherry. No O3 exposure-plant response relationships were found with any species tested
12      in GRSM (Chappelka et al., 1999b); but, when the data were combined for both parks, a
13      significant correlation r = 0.72) with black  cherry was found for both SUM06 and W126, and
14      injury was the greatest (r = 0.87) at the higher elevations (Chappelka et al.,  1999a).
15           Based on a study in which visible symptoms of O3 injury were characterized for large,
16      mature yellow-poplar and black cherry trees in GRSM (Chappelka et al.,  1999a), Somers et al.
17      (1998) compared radial growth differences among trees classified as sensitive or non-sensitive
18      based on the severity of visible foliar  injury observed over a 3-year period (1991-1993).
19      Significantly more radial growth was  observed over both a 5- and a 10-year period for the non-
20      sensitive compared to the sensitive trees. No significant relationship was found for black cherry
21      tree growth.
22           Vollenweider et al.  (2003), using data collected from continuous forest inventory (CFI)
23      plots across Massachusetts, compared growth rates among either symptomatic or asymptomatic
24      mature black cherry trees. Of the 120 trees sampled in 1996, 47% exhibited visible foliar injury.
25      Using CFI data, growth rates were compared over a 31-year period.  The growth rates for
26      symptomatic trees were reduced by 28% compared with the asymptomatic trees.
27           Because these studies (Somers et al.,  1998; Vollenweider et al., 2003) were not controlled
28      studies and used a small sample of trees, they cannot validly be used to characterize cause and
29      effects related to the visible symptoms and radial growth they  describe. However, the  results
30      indicate the possibility that O3 is correlated with growth losses in some sensitive genotypes,
        January 2005                              9-20        DRAFT-DO NOT QUOTE OR CITE

-------
 1      illustrating the potential usefulness of this visible O3 injury methodology in assessing effects on
 2      the growth rates of mature deciduous trees.
 3
 4           Cultivar Comparisons
 5           The idea of using cultivars or isogenic lines of crop species that differed in O3 sensitivity as
 6      sentinels to determine the ambient effects of O3 in the field was presented in the 1996 O3 AQCD
 7      (U.S. Environmental Protection Agency, 1996).  The rationale was that comparing the ratio of
 8      injury scores or some measure of growth between two different cultivars varying in O3
 9      sensitivity should be indicative of the relative amount of ambient stress to plants at a given
10      location.  A sensitive:resistant ratio close to unity would indicate relatively low O3
11      concentrations and a low ratio higher O3 levels.  Results from locations differing in O3
12      concentrations could be evaluated to develop exposure-response models.  The original protocol
13      was derived using two isogenic lines of white clover (Trifolium repens) differing in O3
14      sensitivity (Heagle et al., 1994b; 1995).
15           This white clover model system has been used in several multi-location studies in the U.S.
16      (Heagle and Stefanski, 2000) and Europe (Ball et al., 2000; Bermejo et al., 2002; Mills et al.,
17      2000).  Heagle and Stefanski (2000) compared results from eight sites over a 2-year period with
18      various exposure indices (SUMOO, SUM06, W126 and others) to determine a best-fit regression.
19      They found that most of the indices preformed similarly. The highest r2 values (0.87-0.93) were
20      obtained using only the later harvests and a 6 h d"1 index (1000-1600 h).  Similar multiple-
21      comparison studies conducted in Europe using the AOT40 index (Ball et al., 2000; Mills et al.,
22      2000) yielded poorer r2 values. Factors such as air temperature, NOX (high levels at some sites)
23      and lower O3 concentrations in Europe were suggested to account in part for the differences
24      between U.S.  and Europe study results. Bermejo et al. (2002), in a study in Spain, improved the
25      model by comparing the biomass ratio of these white clover isolines to measures of O3 uptake
26      (flux) rather than an exposure index (AOT40). Together, these studies indicate that systems
27      such as the white clover model can help reveal O3 exposure-response relationships that can
28      provide valuable information regarding ambient O3  conditions in a given location.  Table 9-3
29      lists the advantages and disadvantages of the use of cultivar comparisons in assessing O3 effects
30      of pi ants.
31

        January 2005                              9-21         DRAFT-DO NOT QUOTE OR CITE

-------
               Table 9-3.  Advantages and Disadvantages of Cultivar Comparisons Used in
                                   Assessment of O3 Effects on Plants
        Advantages
        No chambers required. Plants exposed to ambient conditions of O3, light, temperature, etc.
        Relatively inexpensive.  Equipment needs are minimal. No "chamber effects"
        A high degree of replication possible both within and among locations
        Can conduct studies "in situ"
        Disadvantages
        Preliminary tests to insure sensitivity and growth patterns of genotypes used are consistent
        are needed
        Need measurements of ambient O3 and other meteorological variables (temp, rainfall, etc)
        Have to ensure cultural practices (soil, irrigation, fertilization, etc.) are similar among sites
        Closely monitor plants for presence of other factors that may cause a misinterpretation of results
 1          Dendrochronological techniques
 2          It has been difficult to determine whether O3 significantly affects tree growth and
 3     productivity in the field, because O3 concentrations are omnipresent and tree response to this
 4     pollutant is altered by many factors. The use of dendrochronological techniques to answer
 5     questions regarding ambient O3 effects on forest growth and ecosystem function has recently
 6     emerged as a very useful biomonitoring methodology (Cook, 1990; McLaughlin et al., 2002).
 7     The technique is useful when either instrument or passive O3 monitoring methods are used to
 8     determine ambient O3 conditions.
 9          Initial experiments were primarily correlative in nature and attempted to relate symptoms
10     of visible injury with growth losses as revealed by tree ring analysis (Arbaugh et al., 1998;
11     Benoit et al.,  1983; Peterson et al., 1995; Somers et al., 1998; Swank and Vose,  1990). These
12     studies evaluated radial growth patterns determined by cores removed from trees in the presence
13     or absence of overt O3 injury symptoms.
14          The method has also been adapted to better understand forest ecosystem function
15     (McLaughlin and Downing, 1996; McLaughlin and Downing, 1995; Bartholomay et al.,  1997;
16     McLaughlin et al., 2003).  The response of mature loblolly pine (Pinus Taedd) growing in

       January 2005                              9-22        DRAFT-DO NOT QUOTE OR CITE

-------
 1      eastern Tennessee to ambient O3 and moisture stress was evaluated by McLaughlin and Downing
 2      (1996, 1995) over a 6-year period (1988 to 1983). They made radial growth measurements
 3      from 12 to 37 times per year using dendrometer bands and determined relationships between O3,
 4      moisture stress, and radial growth. Exposures to O3 concentrations > 0.04 |_il L/1 with high
 5      temperatures and low soil moisture resulted in short-term depression in radial growth.
 6      Reductions in growth were estimated to vary from 0 to 15% per year and averaged
 7      approximately 5% per year.
 8           Bartholomay et al. (1997) examined white pine (Pinus strobus) radial growth in eight
 9      stands throughout Acadia National Park, Maine over a 10-year period from  1983 to 1992.  They
10      related growth rates to several factors, including O3 concentration. Ozone levels were negatively
11      correlated with radial growth in seven of the eight stands.  Site characteristics were important in
12      the relationship:  stands growing on shallow, poorly drained soils were most sensitive to O3 in
13      the late portion of the growing  season, possibly due to premature senescence of foliage.
14      However, litterfall measurements were not reported.  Trees growing on better sites were more
15      sensitive to O3 during the entire growing season, indicating the possibility of high O3 uptake
16      rates throughout the growing season. Although these field studies (Bartholomay et al., 1997;
17      McLaughlin and Downing, 1996; McLaughlin and Downing, 1995) did not  compare the firm  O3
18      effects on the two pine species, they indicate that potential interactions exist among O3 and other
19      climatic and edaphic factors, such as temperature and soil moisture.
20           Using both automated and manual dendrometer bands, McLaughlin et al.  (2002) examined
21      the growth response of yellow poplar (Liriodendron tulipifera) trees recently released from
22      competition.  In addition to measuring growth, sap flow measurements were conducted and soil
23      moisture was measured in the vicinity of the trees. However, O3 concentrations were low and
24      there were no negative O3 effects observed in this 1-year study. Advantages and disadvantages
25      of dendrochronology techniques for evaluating whole-tree physiological responses for individual
26      trees and forest stands are listed in Table 9-4.
27           The use and evolution of various dendrochronological methods in the field of air pollution
28      effects research is reviewed in detail by McLaughlin et al. (2002).  Automated dendrometer
29      bands provide a powerful tool for measuring radial growth responses of trees on an hourly or
30      daily basis. Diurnal patterns of growth can be related to water use and O3 concentrations using
31

        January 2005                              9-23        DRAFT-DO NOT QUOTE OR CITE

-------
          Table 9-4. Advantages and Disadvantages of Various Dendrochronological Techniques
                               Used in Assessment of O3 Effects on Plants
         Advantages
         Provide information regarding growth effects under ambient conditions
         Good historical information regarding O3 effects
         Can provide data on daily, and seasonal growth and O3 patterns and relate these to
         physiological function
         Provide information on forest function related to ambient O3 concentrations
         Can link data with process level growth models
         Disadvantages
         Results are generally correlative in nature with no true control
         Need background O3 and meteorological data (historical records)
         Need to account for other factors such as competition, in analyzing data
         Individuals need to be trained in counting growth rings
         Replication can be difficult (expensive and technological limitations)
         Complicated statistical analyses are sometimes required
         Can be expensive, especially if using automated growth (dendrometer) bands
 1     time-series analyses. The major drawbacks of the method are that it is expensive and time
 2     consuming.
 3
 4     9.2.3.5 Calibrated Passive Monitors
 5           Many studies have used passive monitors in the mapping of ambient O3 concentrations,
 6     especially in remote areas (Cox and Malcolm, 1999; Grosjean et al., 1995; Krupa et al., 2001).
 7     Since they are cumulative recording devices, they cannot reveal short-term variations in O3
 8     concentration but only the total exposure over a given interval, usually 7 days.  Thus, they
 9     produce a measurement that resembles the instrumentally derived exposure SUMOO index.
10           Runeckles and Bowen (2000) used the ZAPS  system described in Section 9.2.2.3 to subject
11     both crops and passive monitors (Williams, 1994) to a range of exposures. Passive monitors
12     were also exposed at 16 agricultural field sites along a transect through the Fraser Valley, British

       January 2005                               9-24        DRAFT-DO NOT QUOTE OR CITE

-------
 1      Columbia, Canada. Most field sites were downwind of the Greater Vancouver metropolitan
 2      area. All passive monitors were replaced at weekly intervals and the data from those in the
 3      ZAPS plots were "calibrated" to crop responses by means of Weibull exposure-response
 4      functions. Since the meteorological conditions throughout the valley were reasonably consistent
 5      from site to site, the use of these functions with data from the network passive monitors as inputs
 6      permitted the estimation of crop losses at the network sites.  The overall method was thus a
 7      hybrid of several methodologies.
 8           Although based on a single study, the use of passive monitors has potential for assessing
 9      crop losses at sites removed from locations with known ambient O3 concentrations.  Provided
10      that the network and calibration sites have similar meteorological conditions, the method yields
11      crop loss estimates that are responses to local ambient O3 levels as influenced by local
12      meteorological conditions
13
14      9.2.4  Numerical/Statistical Methodologies
15           Proper experimental design strategies including replication, randomization, and
16      experimental protocols are paramount in O3-effects research. These have been discussed in detail
17      in previous O3 AQCDs (U.S. Environmental Protection Agency, 1986,  1996), as have the
18      different statistical analytical procedures used to determine the probable significance of results.
19      However, new investigative approaches have demanded the adoption of new analytical methods.
20      For example, the use of dendrochronological techniques has led to the use of time-series analysis
21      (McLaughlin et al., 2003) and linear aggregate models (Cook, 1990), as reviewed by
22      McLaughlin et al. (2002).
23           In spite of the rigors of the analyses, many differences occur in the published literature for
24      almost any plant response to O3 stress. Differences inevitably result from different researchers
25      studying different locations, using different experimental methodologies and genetically
26      different plant material even when using a common species.  The techniques of meta-analysis
27      can be used to consolidate and extract a summary of significant responses from a selection of
28      such data.
29           Despite the differences in responses in the 53 primary studies used, a recent meta-analysis
30      by Morgan et al. (2003) of the effects of O3 on photosynthesis,  growth, and yield of soybean
31      (Glycine max) showed "overwhelming evidence for a significant decrease in photosynthesis, dry

        January 2005                              9-25        DRAFT-DO NOT QUOTE OR CITE

-------
 1      matter production and yield ... across all the reported studies on effects of chronic O3 treatment."
 2      The meta-analysis defined O3 stress exposure to -70 ppb for at least 7 days and found average
 3      shoot biomass and seed yield decreases of 34% and 24%, respectively. Furthermore, although
 4      other stress factors such as drought and UV-B did not affect the O3 responses, elevated CO2 was
 5      found to significantly decrease O3-induced losses.
 6           The meta-analysis method clearly has the potential to consolidate and refine the
 7      quantitative exposure-response models for many species. The majority of the reported growth
 8      and physiological responses related to O3 stress are for individual plants, primarily in various
 9      types of exposure chambers.  It is difficult to extrapolate these responses to stand/community,
10      ecosystem, or region-wide assessments, particularly in view of the importance of the significant
11      interactions that may occur between plant responses O3 and other environmental stresses. Along
12      with the shift in effects research to a more ecological approach, these concerns necessitate a
13      move from simple regression analysis to more complex mathematical approaches to handle a
14      wider array of independent input variables than O3 exposure alone. Other independent input
15      variables that must be accounted for include air and soil temperatures, soil moisture, relative
16      humidity, wind speed, and, particularly in the case of natural systems, biotic factors such as pests
17      and pathogens, plant density/spacing, and measures of plant competition.
18           Artificial neural network (ANN) methodology was used by Balls et al. (1995) for
19      "unraveling the complex interactions between microclimate, ozone dose, and ozone injury in
20      clover" and in the study with the protectant chemical EDU, discussed in Section 9.2.3.3 (Ball
21      et al., 1998). The multi-factor model for predicting the effects of ambient O3 on clover
22      developed by Mills et al. (2000) utilized both ANN and multiple linear regression methods.
23           Models incorporating ANNs are of the "regression" type (Luxmoore, 1988) in contrast to
24      "mechanistic" or "phenomenological" models which have wider applicability. Process-level
25      models of either type have been developed at the organelle, individual plant (Constable and
26      Taylor, Jr., 1997; Weinstein et al., 1998), canopy (Amthor et al., 1994), and stand level (Ollinger
27      et al., 1997; Weinstein et al., 2001) and provide estimates of the rate of change of response
28      variables as affected by O3 over time.  However, as pointed out in the 1996 Criteria Document
29      (U.S. Environmental Protection Agency,  1996),  mechanistic process models lack the precision of
30      regression models as well as their ability to estimate the likelihoods of responses. In their
31      extensive reviews, Kickert and Krupa (1991) and Kickert et al. (1999) summarized the

        January 2005                              9-26        DRAFT-DO NOT QUOTE OR CITE

-------
 1     advantages and shortcomings of many different models and make the important point that most
 2     of the models that have been described provide consequence assessments that quantify the
 3     magnitudes of effects, but not risk assessments that quantify the likelihoods of such effects.
 4     Descriptions of several specific models are provided in other sections of this criteria document,
 5     and advantages and disadvantages of modeling techniques used in assessing O3 effects on plants
 6     are summarized in Table 9-5.
 7
         Table 9-5.  Advantages and Disadvantages of Modeling Techniques Used in Assessment
                                        of Oa Effects on Plants
        Advantages
           • Provides an understanding of cause-effect relationships over time
        Disadvantages
           • Have to make assumptions based on scarcity of data
           • Most models are very complex and difficult to understand
           • Need to be evaluated for predictive validity
 1     9.2.5  Improved Methods for Defining Exposure
 2          Ambient air quality is defined in terms of the measured O3 concentrations in the air at some
 3     standard elevation above ground level. Compilations of such concentration data have long been
 4     used as surrogates of the exposures to which plants are subjected.  However, as long ago as
 5     1965, field research provided evidence that plant response was a function, not of ambient O3
 6     concentration/>er se, but of the estimated flux of O3 to the plant canopy (Mukammal, 1965).
 7     Subsequently, Runeckles (1974) introduced the term "effective dose" to define that part of the
 8     ambient exposure that was taken up by a plant. Fowler and Cape (1982) later referred it as
 9     "pollutant applied dose" (PAD), defined as the product of concentration, time and stomatal
10     (or canopy) conductance, with units g m"2.  Such estimates of O3 uptake or flux provide a more
11     biologically relevant description of exposure than the simple product of concentration and time
       January 2005                             9-27        DRAFT-DO NOT QUOTE OR CITE

-------
 1      alone, and formed the basis of Reich's 1983 "unifying theory" of plant response to O3 (Reich,
 2      1983).
 3           However, it was not until the early 1990s that the inherent advantages of using O3 flux
 4      rather than O3 concentration as a basis for determining response effects began to be widely
 5      accepted, as demonstrated by the subsequent increase in publications involving flux
 6      measurements and  modeling (e.g., Fuhrer et al., 1997; Griinhage and Jager, 2003; Griinhage
 7      et al., 1993, 1997; Massman et al., 2000; Musselman and Massman, 1999; Pleijel, 1998). A key
 8      requirement for flux determination is the measurement of stomatal or canopy conductances,
 9      using established porometer/cuvette techniques or eddy correlation methods. The usefulness and
10      relevance of flux as a measure of exposure are discussed in detail in Section 9.5.
11           Efforts to  develop models of O3 deposition and stomatal uptake are currently under way
12      with a view to providing improved assessments of the risks to vegetation across Europe
13      (Emberson et al., 2000; Simpson et al., 2001; Simpson et al., 2003).
14
15
16      9.3  SPECIES RESPONSE/MODE-OF-ACTION
17      9.3.1  Introduction
18           The evaluation of O3 risk to vegetation requires fundamental  understanding of both the
19      functioning of the vegetation and how external environmental influences can alter that function.
20      For biological organisms subjected to atmospheric O3, those alterations can be complex and
21      multiple. In addition, biological organisms have plasticity to external interactions due to their
22      complex internal, self-correcting systems, making the task of identifying their "correct"
23      functioning difficult.  This section emphasizes reactions of O3 with the cell and tissue rather than
24      the whole plant to describe the fundamental mechanisms known to govern the response of the
25      plant to O3 exposure.
26           The many regulatory systems contained in leaves change both as a function of leaf
27      development and in response to various environmental stresses. Leaves function as the major
28      regulators of anatomical and morphological development of the shoot and control the allocation
29      of carbohydrates to the whole plant (Dickson and Isebrands,  1991). This section discusses the
30      movement  of O3 into  plant leaves and their biochemical  and physiological responses to O3.
       January 2005                             9-28        DRAFT-DO NOT QUOTE OR CITE

-------
 1           The 1996 criteria document (U.S. Environmental Protection Agency, 1996) assessed the
 2      information available at that time concerning the biochemical and physiological responses to the
 3      movement of O3 into plant leaves. This information continues to be valid. Ozone uptake in a
 4      plant canopy is a complex process involving adsorption to surfaces (leaves, stems, and soil) and
 5      absorption into leaves (Figure 9-1). However, the initial biochemical changes that result within
 6      leaf cells after the entry of O3 and how these changes interact to produce plant responses remain
 7      unclear. The response of vascular plants to O3 may be viewed as the culmination of a sequence
 8      of physical, biochemical, and physiological events. Only the O3 that diffuses into a plant
 9      through the stomata (which exert some control on O3 uptake) to the active sites within a leaf
10      impairs plant processes or performance. An effect will occur only if sufficient amounts of O3
11      reach sensitive cellular sites that are subject to the various physiological and biochemical
12      controls within the leaf cells. Ozone injury will not occur if (1) the rate and amount of O3 uptake
13      is small enough for the plant to detoxify or metabolize O3 or its metabolites, or if (2) the plant is
14      able to repair or compensate for the O3 impacts (Tingey and Taylor, 1982; U.S. Environmental
15      Protection Agency, 1996).  Therefore, a precondition for O3 to affect plant function is that it
16      must enter the stomata and be absorbed into the  water lining the mesophyll cell walls. The
17      response of each plant is determined by the amount of O3 entering the leaves, which varies from
18      leaf to leaf.
19           Some potentially significant processes have been investigated since the 1996 criteria
20      document, especially detoxification and compensatory processes.  The role of detoxification in
21      providing a level of resistance to O3 has been investigated; however, it is still not clear as to what
22      extent detoxification can protect against O3 injury. Data are needed especially on the potential
23      rates of antioxidant production and on the  subcellular localization of the antioxidants. Potential
24      rates of antioxidant production are needed to assess whether they are sufficient to detoxify the O3
25      as it enters the cell.  The subcellular location(s) are needed to assess whether the antioxidants are
26      in cell wall or plasmalemma locations that permit contact with the O3 before it has a chance to
27      damage subcellular systems.  Various forms of compensation, especially the stimulation of new
28      leaf production and of higher photosynthetic performance of new leaves,  have been reported.
29      Although these processes divert resources  away from other sinks, these forms of compensation
30      may counteract the reduction in canopy carbon fixation caused by O3.  The quantitative
31      importance of these processes requires investigation.

        January 2005                               9-29        DRAFT-DO NOT QUOTE OR CITE

-------
Atmospheric Processe
i


	 [Ozone in the | f Canopy |
5 "" 1 Atmosphere 1 1 Boundary Layer 1
r
Canopy Processes
i
r
Leaf Processes/
Ozone Uptake S
^
r
Leaf Processes/
Mode of Action
I
Plant Response
\

Ecosystem Responses
•*^^
"** *«N
k
\
\
\
\
\
\
\
\
\
\
\


~D>
^

K Canopy
Conductance
r :-' -
/^Ozone in N
Vegetation
\^ Canopy J
[>
T

^,
J y
•• • ji p '- ' •-
*, \&f -,',•:'---
~>^!
f
f Leaf Boundary Layer J
S
^^ Leaf
^\_ Conductance
V7
r C Rtn
(f ^ 1 Cond
Ozone Absorption V
/ r-
?^
^^L
matal |
uctance I
1 into Foliage 1
      Figure 9-1.  Ozone uptake from the atmosphere.  Ozone moves from the atmosphere
                  above the canopy boundary layer into the canopy primarily by turbulent
                  air flow. Canopy conductance, controlled by the complexity of the canopy
                  architecture, is a measure of the ease with which gases move into the canopy.
                  Within the canopy, O3 is adsorbed onto surfaces as well as being absorbed
                  into the foliage. Foliage absorption is controlled by two conductances, leaf
                  boundary layer and stomatal, which together determine leaf conductance.
                  The solid black arrows denote O3 flow; dotted arrows indicate processes
                  affecting uptake or response to O3. Boxes at the left with double borders
                  are those processes described in the figure.
1          As a result of the research since the 1996 criteria document (U.S. Environmental Protection

2     Agency, 1996), the way in which O3 exposure reduces photosynthesis, especially its effects on

3     the central carboxylating enzyme, Rubisco (ribulose-l,6-P2-carboxylase/oxygenasel), is better

4     understood. The rate of leaf senescence has been shown to increase as a function of increasing

5     O3 exposure. The mechanism of the increased senescence is not known, and, hence, it deserves
6     further study.
      January 2005
9-30
DRAFT-DO NOT QUOTE OR CITE

-------
 1           Finally, the role that changes in allocation of resources play in plant response to O3 is now
 2      better understood. Most studies have shown that O3 decreases allocation of photosynthate to
 3      roots. In some cases, allocation to leaf production has increased.  Whether these changes are
 4      driven entirely by changes in carbohydrate availability or are controlled by other factors (e.g.,
 5      hormones) is not known. Physiological effects within the leaves inhibit photosynthesis, alter the
 6      assimilation of photosynthate and shift its allocation patterns, and can lead to reduced biomass
 7      production, growth, and yield (U.S. Environmental Protection Agency, 1986, 1996).
 8           The major problem facing researchers trying to predict long-term O3 effects on plants is
 9      determining how plants integrate the responses to O3 exposures into the overall naturally
10      occurring responses to environmental stressors.  Little is now known about how plant responses
11      to O3 exposures change with increasing age and size, but this information is crucial to predicting
12      the long-term consequence of O3 exposure in forested ecosystems.
13           This section focuses on reactions of O3 within cells and cellular tissue, in order to explain
14      known mechanisms that govern plant responses.  The processes that occur at cell and tissue
15      levels within the leaf will be divided into several steps beginning with O3 uptake and its initial
16      chemical transformations into a series of currently unknown, but suspected toxic, chemicals
17      (Figure 9-2). The discussion will then focus sequentially upon various cell regions, their general
18      physiology, and the changes that may occur within a plant after O3 exposure. This is important
19      because the varying responses of the different plant species in a community ultimately lead to
20      an ecosystem response. Finally, a general summary is presented that discusses the known or
21      suspected changes that occur within the whole plant.
22
23      9.3.2  Mechanisms of Ozone-Induced Plant Alterations
24           Plant adaptations for surviving O3 stress include exclusion or tolerance of it or its products
25      (Levitt, 1972; Tingey and Taylor, 1982).  Ozone may be excluded from tissues or cells via
26      stomatal closure, by extracellular oxidants, or by membrane impermeability to O3 or its products.
27      Past investigations of O3 injury have indicated that  physiological and metabolic changes occur
28      (see Heath,  1998; Reddy et al.,  1993; Harris and Bailey-Serres, 1994).  Many of these changes
29      are likely initiated via gene expression. During the last decade, our understanding of the cellular
30      processes within plants has increased.  Although the fundamental hypotheses concerning
        January 2005                              9-31        DRAFT-DO NOT QUOTE OR CITE

-------
                     Atmospheric Processes
                    I Canopy Processes |
                           X
                     I Leaf Processes | J
                           x
                I Plant Response Processes |
                          1
                    | Community Processes
                           4
                    I Ecosystem Processes
                                                         Plasma Membrane
Second Sites:
Signaling

Second Sites:
Metabolism
                                                           General Physiology:
                                                            Cell Metabolism
                                                          General Physiology:
                                                             Whole Plant
      Figure 9-2. Absorption and transformation of O3 within the leaf. The varied processes are
                  broken down in to smaller mechanistic steps that lead from uptake of
                  atmospheric O3 into the alterations which may occur within the individual
                  plant. Each plant responds to the O3 level and therefore interacts with the
                  total ecological setting to generate an ecosystem response due to the O3.
1
2
3
4
5
O3-induced changes in physiology have not changed, a more complete development of the
theories is now nearing possibility.

9.3.2.1  Changes in Metabolic Processes: Current Theories
     The current hypotheses regarding the biochemical response to O3 fumigation revolve about
injury and its prevention.  These are well discussed by Pell et al. (1997) and are listed below in
no order of importance. Although they are listed separately, some may be interlinked and related
to each other.
      January 2005
                                          9-32
DRAFT-DO NOT QUOTE OR CITE

-------
 1           (1)   Membrane Dysfunction. The membrane is altered by O3, principally via protein
                  changes not involving the lipid portions of the membrane (except at extremely high
                  levels of O3).  These alterations involve increased permeability with perhaps lessened
                  selectivity, declines in active transport, and changes in the trigger mechanisms of
                  signal transduction pathways such that the signals are no longer suitable for the state
                  of the cell. The cellular pools and transport systems of Ca2+/K+/H+ are the primary
                  suspects.

 2           (2)   Antioxidant Protectants. Varied antioxidants (both as metabolites and enzyme
                  systems) can eliminate the oxidant or its products, if present at time of fumigation
                  and in the  sufficient abundance.  However, oxidant entry that occurs rapidly  can
                  overwhelm the antioxidant response.

 3           (3)   Stress Ethylene Interactions. Visible injury is caused by the interaction of O3 with
                  stress-induced ethylene, either by direct chemical transformation to a toxic product or
                  by alteration of the biochemical relations at the ethylene binding site.

 4           (4)   Impairment of Photosynthesis.  A product of O3 (and less probably, O3 itself) enters
                  the cell, causing a decline in the mRNA for Rubisco (especially the message RNA
                  species ofrbcS and rbcL) such that Rubisco levels slowly decline within the
                  chloroplast, leading to a lowered rate of CO2 fixation and productivity. This process
                  is very similar to early senescence and may be linked to general senescence.
                  Alternatively, a false  signal  is generated at the cell membrane which lowers the
                  transcription of DNA to mRNA. Ozone alters the normal ionic and water relations of
                  guard cells and subsidiary cells, causing the stomata to close and limit CO2 fixation.
                  In any case, the response of the stomata to the current environment does not  promote
                  efficient photosynthesis.

 5           (5)   Translocation Disruption. One of the biochemical systems most sensitive to O3
                  exposure is the translocation of sugars, such that even a mild exposure inhibits the
                  translocation of carbohydrate (Grantz and Farrar, 1999, 2000).

 6           (6)   General Impairment/Disruption of Varied Pathways of Metabolism. This is the
                  oldest and most vague concept of how O3 alters metabolism.  It is based upon early
                  work in which what enzymes and metabolites could be assayed were.  Thus,  these
                  results were based upon what could be done, rather than a coherent hypothesis. The
                  best examples are listed in Dugger and Ting (1970).

 7           The latter two theories can be restated as a loss of productivity with three possible

 8      somewhat-independent causes — (a) a reduced production of the basic building blocks of growth

 9      and, hence, a slowing of growth in at least one organ; (b) a reduced ability to reproduce, leading

10      to a decreased production of viable seeds or of fruits and nuts; and (c) a decreased ability to

11      mount a defense against pathogens or insects, leading to weaker plants, which are more  liable to

12      be overcome by other stresses. It is important to separate out effects that may be detrimental or
       January 2005                               9-33        DRAFT-DO NOT QUOTE OR CITE

-------
 1      disfiguring, such as the production of visible injury, but which have not been shown to lead
 2      directly to a loss of productivity due to possible compensation by the remaining tissue. The loss
 3      of some localized regions of photosynthesis tissue has not been shown to lead directly to a loss
 4      of productivity.
 5
 6      9.3.2.2  Modifications of Plant Physiological Processes
 7           The discussion that follows will focus on physiological processes; the species used to
 8      develop an understanding of these processes are relatively unimportant. The study of any plant
 9      that can help increase the biological understanding of the response of plants to O3 exposures is of
10      use, regardless of how sensitive it is to O3.  Therefore, Arabidopsis, whose physiology and
11      genome continue to be studied and described by a large number of scientists is an appropriate
12      plant for studying O3 injury.  Though the responses of mature trees and understory plants are
13      critical to understanding plant interactions at an ecosystem level, the time required for trees to
14      reach maturity makes using them to study biological mechanisms a poor choice.
15           The high levels of O3 used for some investigations do not automatically invalidate the
16      results  obtained in those studies. Typically when a new hypothesis is being investigated,
17      extreme levels of the toxicant are used  to clearly determine its effects.  The older studies that
18      used concentrations as high 1 ppm, an extreme level, helped to define current studies. Later
19      experiments have used concentrations nearer ambient levels.  Many of the current studies on
20      physiology use exposures between 0.15 and 0.25 ppm, which though higher than ambient levels
21      in some areas of the country, bypass confounding changes but allow for rapid experiments.
22           Three forms of air pollutant-induced injury patterns currently are known to exist: (a) acute
23      stress, generated by high atmospheric concentrations of pollutants for short periods of time;
24      (b) chronic stress, generated by lower concentrations of pollutants for long periods of time; and
25      (c) accelerated senescence, generated by very low concentrations of pollutants for very long
26      periods of time.  At higher levels, distinct visible injury generally occurs due to cellular and
27      tissue death of regions of leaf mesophyll cells. This leads to  a decline in the total area of
28      metabolically active tissue, with consequent loss of membrane integrity, loss of metabolites into
29      the extracellular tissue space, and formation of oxidative products. When no visible injury is
30      observed, lowered rates of photosynthesis or productivity are often used to index injury. Under
31      these conditions, metabolism is altered and the pool sizes of many metabolites are changed.

        January 2005                              9-34        DRAFT-DO NOT QUOTE OR CITE

-------
 1      More importantly, the altered biochemical states within the tissue lead to the inability of the
 2      plant to respond properly to existing environmental conditions and to other stressors (Manning
 3      and Keane, 1988; Heath, 1988; 1994b; Koziol and Whatley, 1984; Schulte-Hosted et al., 1988).
 4
 5      9.3.3   Ozone Uptake by Leaves
 6          Plants respond to O3 similarly to other stressors on several levels:  exclusion, tolerance, and
 7      repair (Levitt, 1972; Tingey and Taylor, 1982). The response mechanism depends upon the O3
 8      concentration, environmental conditions, and the developmental and metabolic state of the plant
 9      (Guzy and Heath, 1993). These responses are detrimental to plant productivity because they cost
10      the plant metabolic resources. In some cases, the stomata close under the O3 exposure,
11      excluding the pollutant from the leaf interior and preventing injury. However, if this happens
12      too often, CO2 fixation is also inhibited and plant productivity suffers.
13          Atmospheric O3 does  not cause injury, but rather it is the O3 that enters the plant that
14      causes an effect (Tingey and Taylor, 1982).  Three well-defined, sequential processes control the
15      movement of O3 from the atmosphere into the sites of action within the leaf and must occur to
16      trigger O3 stress (Heath, 1980).  The processes are: (1) entry of O3 into the leaf, (2) reactions of
17      O3 and its possible reaction product(s) in the water phase at cell surfaces, and (3) movement of
18      O3 reaction product(s) into the cell with enzymatic or chemical transformation of those products
19      in the cell.
20
21          Process 1. Entry ofO3 into the leaf. Often incorrectly, the external concentration of O3 is
22      used to give an indication of "dose"  (Heath, 1994a). Ozone-induced changes on the plants
23      cuticle are minimal, and O3 does not penetrate the cuticle (Kerstiens and Lendzian, 1989) to
24      cause an effect.  As O3 has no easily measured isotope, virtually no measurements have been
25      done on an actual dose of O3, i.e., the amount of O3 which reacts with individual biochemicals in
26      the leaf.  Yet the measurement of dose will be the amount of O3 expected to penetrate into the
27      tissue through the stomata.  Dose is expressed as a rate of delivery to a surface area (mol/m2 s"1).
28      Whether dose or total accumulation (mole/m2, rate integrated over exposure time) is most critical
29      for the development of injury remains a major question.
30          Ozone uptake includes gaseous diffusion through the leaf boundary layer and stomata into
31      the substomatal cavity (Figure 9-3).  Although the movement of pollutants through a boundary

        January 2005                              9-3 5       DRAFT-DO NOT QUOTE OR CITE

-------
Bulk Air Z^>
Boundary L-hv
Layer \~^
Stomatal 1— 1\
Aperture V~r
Sub-
stomatal
Cavity
Wind Speed Leaf
Canopy Temperature
Structure
       Figure 9-3.  The uptake of O3 into the leaf. Each of the individual concentration
                    layers of O3 represents a different process of movement and of plant/
                    microenvironmental interaction.  This figure leads into Table 1, in which
                    the amounts of O3 along the pathway are calculated.
 1     layer into the stomata region is known to be important, and even rate limiting in many cases of
 2     low wind velocity, its description has been defined from aeronautical concepts and usually
 3     relates to smooth surfaces that are not typical of leaf-surf ace morphology; however, it is nearly
 4     the only treatment available (Gates, 1968). Once through the boundary layer, the gas must enter
 5     the leaf through the stomata. The entry of gases into a leaf is dependent upon the physical and
 6     chemical processes of gas phase and surfaces and is a  well defined path that approximately
 7     follows a linear flux law of:
 8
                                         j = g(C0-Ci)                                   (9-1)

 9
10     where the flux,y, into the internal space of a leaf is related to the conductance, g, through the
11     boundary layer and stomata and the gradient of concentration of gas from the outside, C0,
12     inwards,  C;. This formulation has been used for years for both water and CO2 (Figure 9-4),  and
13     for regions of varied CO2 concentration that correspond to C0 (CO2 of the atmospheric air, below
14     the leaf proper) and Q (CO2 near the leafs spongy mesophyll cells) (Farquar and Sharkey, 1982;
15     Ball, 1987).
16          In the past, the internal concentration of O3 has been assumed to be zero (Laisk et al.,
17     1989), due  to early studies that found that virtually no O3 could pass through a leaf.  That was
18     expected because O3 is extremely reactive with cellular biochemicals.  If the assumption that the
       January 2005                              9-36        DRAFT-DO NOT QUOTE OR CITE

-------
                                            Light
                         Cuticle
                   Epidermis    r
                      Pallisade
                     Mesophyll
                       Spongy
                      Mesophyll
                   Epidermis
                        Cuticle
nnnnnn
            Vascular
             System
                             C0=[C02]-
       Figure 9-4.  The microarchitecture of a dicot leaf. While details among species vary,
                  the general overview remains the same. Light that drives photosynthesis
                  generally falls upon the upper (adaxial) leaf surface. Carbon dioxide and
                  O3 enters through the stomata on the lower (abaxial) leaf surface, while
                  water vapor exits through the stomata (transpiration).
 1     internal concentration zero is correct, then the effective delivery rate for O3 is given as g x C0,
 2     with stomatal conductance being the major regulatory control (Taylor, Jr. et al., 1982; Amiro
 3     et al., 1984). However, a recent study by Moldau and Bichele (2002) indicated that the internal
 4     O3 concentration may not be zero, as previous assumed.  Moldau and Bichele (2002) permitted
 5     leaves ofPhaseolus vulgaris L., which have stomata on both upper and lower leaf surfaces, to
 6     take up O3 at a high rate for 3 to 5 min. Exposure of the lower leaf surface resulted in up to 5%
 7     of the O3 that was taken up to be diffused through the leaf, emerging from the stomata on the
 8     upper surface. This suggested above-zero concentrations of O3 in the intercellular leaf air
 9     spaces. The descriptive calculations and plots of Moldau and Bichele (2002) indicate that the
10     rise in internal O3 level  (for both concentrations of external O3) within the first few minutes of
11     exposure is due to its reaction with an antioxidant, most probably absorbate, within the
12     apoplastic space of the leaf (Figure 9-5). The rate of rise is probably due  to more complete
13     penetration of O3 with a concurrent depletion of the external antioxidant.  The rise peaks at about
14     2 min for 0.88 ppm and 3 min for 0.34 ppm and then falls to a lower level. This may be due to a
       January 2005
              9-37
DRAFT-DO NOT QUOTE OR CITE

-------
                         160
                         140
                         120
o
E
c
0)
o
N
O
                         100
                          80
                          60
         03 Exposure
           D  0.335
           •  0.821
                                                      I
                                                          I
                                           234
                                            Time (minutes)
                                                    40

                                                    38
                                                       g
                                                    36 ^
                         "
                            CD
                         32 §
                            N
                            O
                                                                          30
                         28
       Figure 9-5.  The change in the O3 concentration inside a leaf with time. Data are from O3
                    exposures at two different concentrations.
       Source:  Derived from data in Moldau and Bichele (2002).
 1     replenishment of the antioxidant. The authors saw no injury to the plasmalemma (as measured
 2     by penetration of a dye) and no change in the stomatal conductance for the lower concentration
 3     of O3 (Moldan and Bichele, 2002).  The higher level (0.88 ppm) caused the plasmalemma of the
 4     mesophyll cells to pass a dye and a slight decline in stomatal conductance resulted at about
 5     2.5 minutes.  These data suggest that the antioxidant hypothesis is correct.
 6           Gaseous pollutants flow from the substomatal cavity within the leaf into the cell, through
 7     the cell wall. It is suspected that the internal concentration of the pollutant is not uniform within
 8     the cavity. From within the wall, an equilibrium between the gas and aqueous phase must occur
 9     at the interface where the gaseous species dissolve into the water according to Henry's Law
10     (Heath, 1980, 1987; Wellburn, 1990). It is important to understand exactly how much O3 could
11     move into the tissue of the leaves. Calculations in Table 9-6 give an indication of the amount of
12     O3 which may end up near the surface of cells within the leaf.  The calculation is done for a
13     standard temperature (25 °C), an ambient concentration of O3 (0.10 ppm), and for nonspecific
       January 2005
9-38
                                       DRAFT-DO NOT QUOTE OR CITE

-------
               Table 9-6. The Flow of Ozone into a Leaf and Possible Reactions

 The level of O3 in the atmosphere is chosen to be close to a standard and yet make calculations to other amounts
 easy.  The same concept will be used for all standard parameters for these calculations.
 DESCRIPTION

 The atmospheric level of O3 is given as:

 For an air temperature of:

 The perfect gas law (pV = nRT) is used to convert the O3
 level into standard mks. Further, the volume for a mole of
 gas (V0 = 22400 m3) will be used, from the perfect gas law
 with R= 8.3144

 Thus, the concentration of O3 within the atmosphere is:
                O3ax 10-6-(Ta + 273.18)

          °3           V0 x 273.18

 The stomatal conductance of the gas must be chosen to be
 standard but adjustable. The number should be as large as
 typically measured, but allow for easy conversion, if
 necessary. For a stomatal conductance of:

 The amount of O3 that will penetrate inside the leaf (for a
 typical concentration of nearly zero inside the leaf),  is:
             °3T
                       18  gsvvv
                      48   100
 In terms of amount of water within the leaf we can assume
 that about 85% of the weight is water and the density of
 water is 1 g/ml.  A typical leaf has a wet weight/area:

 Thus, the square surface area of the leaf will translate into
 water space (for concentration of chemicals), as:
              ArT .=
FWL x 0.85
    100
                                                    VALUES

                                                  O3a = 0.1ppm

                                                    Ta = 25 °C
                                                = 4.873 xlo~12 moles/m3
                                                       = 1 cm/s
                                           O3L = 2.984 xlo~14 mol/(m2 s)
                                                 FWL = 30 mg/cm2
                                                                      Ar, = 0.255 L/m2
 The maximum amount of toxic compound that will be
 generated, assuming all the O3 is converted, is given
 below. Here the units of the leaf area weight are
 converted into the mks system and the water space units
 are converted into L, such that the concentrations
 calculated will be in mol/(L hr). The final units assume
 that the O3 is present (and no back reactions occur) for one
 hour (short but typical units of exposure).

           O3 =(O3 /ArL)x3600
                                            O3Lc = 4.21 x 1Q-10 mol/(L hr)
January 2005
                            9-39
DRAFT-DO NOT QUOTE OR CITE

-------
                 Table 9-6 (cont'd). The Flow of Ozone into a Leaf and Possible Reactions
         Thus, the maximum amount of toxic chemicals generated
         per hour in a leaf would be:                                     O3Lc = 0.42 nmol/(L hr)

         Possible errors in these calculations (aside from the input numbers) are: (1) the O3 within the leaf does not react
         uniformly within the leaf space, (2) the O3 within the leaf does not totally  convert to any one species, (3) varied
         products of O3 react leading to innocuous chemicals, and (4) O3 reactions  can be catalytic and generate more by
         radical reaction cycling.
 1      leaves. For example, 0.3 ppm would be the same general numbers but multiplied by 3.
 2      Similarly for more closed stomata, the value of 1.0 cm/s (equivalent to about 400 |imole~2-leaf
 3      area s"1) for a conductance would be reduced and the smaller values would lead to a smaller
 4      amount of O3 moving into the tissue.  Nonetheless, these values give some indication of what
 5      sort of chemical concentration can be expected. Under these conditions, a delivery rate of O3
 6      into the substomatal cavity near the spongy mesophyll tissue of about 0.42 nmol/(L hr) appears
 7      to be reasonable.
 8
 9      Process 2. Ozone diffuses into the leaf air spaces and reacts either with varied biochemical
10      compounds (path 1) which are exposed to the air or is solubilized into the water lining the cell
11      wall of the air spaces (path 2). As shown in Figure  9-6, each reaction has the possibility of
12      transforming O3 into another chemical species (a toxicant) which, in turn, may react with other
13      chemical species and lead to a cascade of reactions.
14           Within the stomata, gases react with the water at the cell's surface and generate new
15      species with the components within the cell wall region. The possible varied pathways are
16      depicted in Figure 9-7.  Although these chemical reaction are poorly understood, some of the
17      fundamentals are known (Heath, 1987; 1988; Wellburn, 1990).  Ozone reacts with organic
18      molecules at the double bonds to form carbonyl groups and, under certain circumstances,
19      generates peroxides, such as hydrogen peroxides (H2O2), superoxide (O2 ) and its protonated
20      form (HO2'), hydroxyl radical (HO'), and peroxy radical (HO2').  Other chemicals present in the
21      water phase can lead to many other oxygenated moities (Figure 9-6). Each of the steps are
22      generally pH dependent (Walcek et al.,1997; Jans and Hoigne, 2000).
23
        January 2005                              9-40        DRAFT-DO NOT QUOTE OR CITE

-------
                  Plasma
   Apoplasm    Membrane
Toxicant1A .

Reaction,*
   Solubilization -
       4
    Reaetion2A
       4
    Toxicant2A. -
                                                             • Reaotion1p
             • Reaotiorijp
                                                             'Reaction1p-
                                                                         Cytoplasm
                              Reaction1c
                                                                         • Reaction2c
                             • Reaction2C-
  Figure 9-6.  Possible transformations of O3 within a leaf.
                         a.
                         Hydroxyl
                         Radical
                         b.
                                          HH
      Superoxide
                                                   t

            ^H202

             Hydrogen
             Peroxide
                                                 HO-      H2O2
                                                 H2O
         ^^
                                                         Peroxyl
                                                         Radical
Figure 9-7.  Possible reactions of O3 within water,  (a) Ozone reacts at the double
             bonds to form carbonyl groups,  (b) Under certain circumstances,
             peroxides are generated.
January 2005
9-41         DRAFT-DO NOT QUOTE OR CITE

-------
 1           Sulfhydryls are particularly easy targets, with the formation of disulfide bridges or sulfones
 2      (Mudd and Kozlowski, 1975).  In water, the reactions become more confusing, but some
 3      products have been described by Heath and Castillo (Heath and Castillo, 1987), such as H2O2,
 4      HO', and O2  (Figure 9-7). Effective detoxification reactions can occur here via antioxidant
 5      metabolites and enzymes, such as ascorbate, glutathione (GSH), and superoxide dismutase
 6      (SOD), if they are present at high enough concentrations (Matters and Scandalios., 1987;
 7      Castillo et al., 1987; Fong and Naider, 2002). If the levels are low, it is believed that stimulation
 8      of their production is a response to O3, albeit a slow one (Harris and Bailey-Serres, 1994).
 9      Certainly it is possible that chemical modification of wall-specific biochemicals (Castillo et al.,
10      1987), such as glucan synthase (Ordin et al., 1969) and diamine oxidase (Peters et al., 1988), is
11      possible.
12           Process 3. Movement of reaction product(s) into the cell and enzymatic or chemical
13      transformations within the cell.  It is believed that the initial site of O3 injury is near and within
14      the plasma membrane. Certainly, membrane functions, such as membrane fluidity (Pauls and
15      Thompson, 1980), permeability (Elkiey and Ormrod, 1979), K+-exchange via ATPase reactions
16      (Dominy and Heath,  1985) and Ca2+ exclusion (Castillo and Heath, 1990), are changed. The
17      similarity of wounding responses (Langebartels et al.,  1991) and O3-induced membrane
18      disruption suggests the induction of normal wound-regulated genes (Mehlhorn  et al., 1991;
19      Sandermann, Jr.,  1998). This implies that O3 can react with components of the cell wall
20      connected to the cytoplasm through the cell wall and membrane by membrane-specific proteins
21      not directly linked to transport.
22           Ozone is soluble in water and once having entered the aqueous phase, it can be rapidly
23      altered to form oxidative products that can diffuse more readily into and through the cell and
24      react with many biochemicals.  Again, the presence of internal antioxidant would be critical to
25      reduce the concentration of most oxidants.  A toxic product of O3 may migrate through the
26      cytoplast to react with photo synthetic processes, or a spurious signal generated at the membrane
27      may affect some control process or signal transduction pathway (Schraudner et al., 1998;
28      Overmyer et  al., 2000, 2003; Rao et al., 2000a, 2002; Sandermann, 2000; Rao and Davis, 2001;
29      Leitao et al.,  2003; Vahala et al., 2003; Booker et al., 2004).
30
31

        January 2005                             9-42        DRAFT-DO NOT QUOTE OR CITE

-------
 1      9.3.3.1  Possible Reactions Within the Leaf
 2           Ozone can react with many compounds within the substomatal cavity of the leaf1 to
 3      produce a variety of oxidizing and toxic chemicals.  Some of the possible reactions which will
 4      generate H2O2, HO', and SO2 , as well as charged O3 intermediates are indicated in Figure 9-8.
 5      Many of these complex reactions have been studied within water solutions through research of
 6      O3 induced water purification and are very dependent upon solutes present with the solutions,
 7      including FT (see von Gunten (Von Gunten, 2003)). An important point is that in alkaline
 8      media, O3 forms H2O2, but in acid media O3 is relatively stable in the absence of free metal ions.
 9           The rates of reaction of O3 with several important compounds, including those with a
10      double bond, the so-called Crigee Mechanism shown in Figure 9-8, can be calculated from the
11      reaction coefficient as given by Atkinson (1990) (Table 9-7). The double bond of the ascorbate
12      molecule is particularly sensitive to O3 attack.  An unstable ozonide product is formed and then
13      accelerates the breakage of the double bond, leading to the formation of two products because of
14      the ring formation of the ascorbate molecule.  These products are relatively unstable and can
15      lead to further reactions not shown in Figure 9-8. The rates of reactions can be calculated
16      (Heath, 1987) from the concentrations of O3 that should occur within the wall regions (Table
17      9-6). At a local concentration of 25 jiM O3, it would take  5000 s (83  min) for all of the O3 to
18      react if there was no further flow of O3.  Clearly, O3 does not react rapidly with the compounds
19      in Table 9-7 and, although some of the products would be  formed through the Crigee Mechanism
20      (see Figure 9-8a), they would be low in concentration2. While other radicals, such  as hydroxyl
21      radical (see Figure 9-8b) can attack double bonds, the products differ. Of particular note for
22      later discussion,  is the reaction of O3 with ascorbate (see Mudd, 1996; Figure 9-8b), which will
23      cleave the double bond in the ring. Unfortunately, little work has been done to characterize
24      possible products within the leaf (but see next section).
25           In a paper discussing the stability and reactivity of O3 in the pulmonary air/tissue
26      boundary, Pryor (1992) calculates that O3 has a half-life of about 7 x  10~8 s in a bilayer.
               lrThe volume of the substomatal cavity (that are within the leaf immediately below the stomata) must be
        regarded as the region in which most O3 reactions occur.  That volume, at a relative humidity of near 100%,
        possesses many diverse surfaces with varied bonding, which could alter the fate of O3.
                For example, hydroxylmethyl hydroperoxide would be expected to be formed by the reaction of O3
        with ethylene and its effects have been tested on peroxidases (Polle and Junkermann, 1994). Unfortunately,
        the concentration of required for inhibition is much higher than would be expected to be formed within the leaf.

        January 2005                               9-43        DRAFT-DO NOT QUOTE OR CITE

-------
    a.
N0
                      Crigee         n

                      Mechanism    / \

              ;=CH2   	»-    9    P

                               H2C  — CH2
             OH-  +   H2C=CH2
                       H2C=CH2
  OH
   \

H2C —



  ON02


H2C —
                                                                   H

                                                                 HC=0


                                                                  0
                                                                   ii

                                                                 HC-OH
    b.
                                          CH(OH)CH O2H
                                          CH(OH)CH O2H
                                              HoO
                                               2u2
                                HO    OH
                                        \
                           O=C          CH(OH)CHO2H


                                CHO , CHO
                             Further Oxidation
Figure 9-8a,b. The Crigee mechanism of O3 attack of a double bond, (a) The typical

             Crigee mechanism is shown in which several reactions paths from the

             initial product is shown,  (b) Typical reaction of ascorbic acid with O3.



Source: Adapted from Mudd (1996).
January 2005
                       9-44
   DRAFT-DO NOT QUOTE OR CITE

-------
        	Table 9-7. Some Rates of Reaction of Ozone With Critical Biochemicals	
         [a] Double bond reactions. The second column is taken from Atkinson (1990) and transformed into Column 3.
         Those rate coefficients are used to calculate the rate of reaction at a concentration of 10 ppm for the organic and
         0.1 ppm for O3 in the air stream within the leaf (localized concentration of about 25 mM, see Table 9-6).
         Compound        x 10~18 cm3/molecules s"1    Rate coefficient (L/mole s"1)     Rate of reaction (M/s)
         Ethane                      1.7                      1.02 x 103                   4.3 x I0~n
         Propene                    11.3                     6.80 x 103                   2.8 x lO"10
         1-butene                    11                      5.91 x 103                   2.5 x 10"10
         trans-2 Butene              200                     1.20 x 105                   5.0 x 10"9
         a-pinene                    85                      5.12 x 104                   2.1 x lO"9
         [b] Possible Oxidative Species. Another possibility is given by the reactions below from Walcek et al. (1997).
         Reactions                                            Rate constants
         (1)  O3 + OH + H2O -> H2O2 + O2 + OH                  kj = 3.67 x 10 mokT1 L s~'
         (2)  O3+ O2 -> HO + 2 O2 + OH                        k2= 1.26 x 109 mole"1 Ls"1
         (3)  O3+ HO2 -> HO + O2 + O2                        k3 = 2.09 x 106 mole"1 L s"1
[c] Possible Concentrations of Other Oxidative Species. Table from Heath (1987). Based upon 100 ppm O3
in gas stream

Species
Superoxide Radical (O2')
Ozone Radical
Protonated O3 radical (HO3')
Concentration
pH7
8.75 x 10~15
4.16 x KT15
1.48 x 1Q-16
(M)
pH 9 Molecules within wall
1 x 10~12 5.5 x KT6 6.3 x 10~4
5 x 1Q-14
1 x IQ-18
        Number of molecules within apoplastic space of (10~12 L) at 0.1 ppm O3.
1      However, the transit time through the lung lining fluid layer is about 2 x 1CT6 s based upon a
2      reasonable estimate for the diffusion of O3. This means that O3 would suffer nearly 29 half-
3      lives3 in passage through the layer, which would reduce O3 to about 3 x 1CT9 of the original
4      concentration — zero for all practical aspects. In the same publication, Pryor points out that any
5      sulfhydryl or ascorbate would interact strongly with O3, further reducing its net concentration.
               3Here a half-life is the time that it takes the reactive species to travel a distance in which it loses 50% of its
       initial concentration. Therefore for a 29 half-life, the concentration has been reduced by 2~29 or about a 10~9 decline.

       January 2005                                 9-45        DRAFT-DO NOT QUOTE OR CITE

-------
 1      The reactivity of cysteine is 109, while the reactivity of tryptophan, methionine, polyunsaturated
 2      fatty acids, and tyrosine is about 2 x 106, and that of phenylalanine is only 103. These numbers
 3      are similar to what has been found for O3 reactivity with amino acids and proteins in aqueous
 4      solutions. In glycophorin (Banerjee and Mudd, 1992) and cytochrome C (Mudd et al., 1996,
 5      1997b) in aqueous solutions, only the methionine was oxidized by O3, producing sulfoxide.
 6      In other proteins lacking methionine, tryptophans were oxidized only if they were in an exposed
 7      position on the surface of the proteins (Mudd et al., 1997b).  Treatment of red blood cell ghosts
 8      with O3, oxidized peripheral proteins of the plasma membrane before it oxidized lipids (Mudd
 9      etal., 1997a).
10
11      9.3.3.2 Toxicants Within the Wall Space
12           While Mehlhorn et al. (1990) is often thought to have shown that free radicals were formed
13      in plant leaves under O3 exposure, careful reading of that paper clearly shows that there was no
14      real evidence of free radicals induced by O3.  Living tissues have many free radical signals,
15      making it difficult to observe changes in free radicals.  Further, the work of Grimes et al. (1983)
16      has also been cited as showing the presence of free radicals in living tissues due to O3 exposure;
17      however, no radical signals were found unless certain organic acids (e.g., caffeic  acid) were
18      added to the tissue with the O3 exposure.  They used the radical trap TMPO (tetramethylphrrolise
19      1-oxide) which reacts with  many types of free radicals to form a stable radical that can be used
20      to "trap" or increase the amount of radical present (see Figure 9-9a).  Ozone would directly react
21      with this trap only if it were bubbled into the solution, not passed over the top of the solution.
22      In the presence of sorbitol or caffeic acid, the trap would indicate the presence of OH radical,
23      which would mean that O3  ->• HO'. Superoxide dismutase, catalase, or EDU had no effect upon
24      this signal, suggesting O2 and H2O2 were not involved in the above sequence. Both O3 and O3
25      plus caffeic acid had no effect upon the protoplasts' intactness or viability.  Thus, 10~5 M HO'
26      and/or 0.30 to 0.40 ppm O3 did not react with the cell membrane. They found no signal in
27      normal cells after subjecting the leaf to O3 and concluded that the radicals were produced via  a
28      concerted mechanism with  the acid. This does not fit with the mechanism postulated by
29      Mehlhorn et al. (1991), which involved a reaction of wound-induced ethylene and O3 at the wall
30      level to generate some free radicals.
31

        January 2005                              9-46       DRAFT-DO NOT QUOTE OR CITE

-------
        a.
        ESR spectra from solution treated
        with O3 at 0.3 pg/g dissolved for
        60 s (1a) control, (1b) with caffeic
        acid; spin trapped: DMPO
        (Grimes et al., 1983)
                                   ESR spectra from bean
                                   (Phaseolus vulgar!s cv. Pinto)
                                   leaves pretreated with spin
                                   trapped:  PEN after a 4-h
                                   fumigation with O3 (0-300
                                   nl/L)(Mehlhornetal., 1990).
         C.

         EPR signal in white light before
         (broken line) and after (dotted
         line) exposure of a bluegrass
         leaf to 1000 ug rrf3 O3 in the
         EPR cavity for 1 h. The solid line
         depicts the difference signal.
         (Runeckles and  Vaartnou, 1997)
       Figure 9-9.   Varied ESR radicals, trapped and not, generated by ozone under somewhat
                    physiological conditions, (a) The generation of a DMPO-trapped radical with
                    caffeic acid in water solution (Grimes et al., 1983).  (b) The generation of a
                    DMPO-trapped radical within bean (Phaseolus vulgaris cv. Pinto) exposed to
                    0.10 ppm O3 for 4 hours. The lower trace is the ESR signal produced with
                    0.3 ppm O3 (Mehlhorn et al., 1990).  (c) The EPR signal produced within a
                    bluegrass leaf exposed to ppm of O3 for 1 h (Runeckles, 1997). Although no
                    trapping agent was used in this experiment, the signal is complex because of
                    various free radicals normally present within the illuminated leaf.
1

2

3

4

5
     The hypothesis that the production of wound-induced ethylene by O3 exposure and its

reaction with O3 would result in the production of radicals was tested by Mehlhorn et al. (1990),

using electron paramagnetic resonance spectroscopy.  After 4 h of 300 ppb, an EPR signal of a

compound was detected which resembled a butonyl radical (Figure 9-9b). Using 70 ppb, the

signal was reduced by about one third of an ethyl radical4, leading to injury.  However, the

spraying of the plant with 1-aminoethoxyvinyl-Gly (AVG), which reduces the production of
              4The reaction would be: O3 + H2C = CH2 -»• varied C-l compounds, due to double bond cleavage, at a
       rate constant of 1.7 x l(T18 cnf/molecule sec = 1.02  x 103 NT1 s"1 (Atkinson R., 1990). This should be compared
       with a reaction of the hydroxyl radical with ethylene, which has a rate constant of 8.52 x 10~12 cm3 molecule"1 s~',
       or 106 x faster.
       January 2005
                                           9-47
DRAFT-DO NOT QUOTE OR CITE

-------
 1      ethylene and visible injury, had no effect upon the EPR signal, suggesting that the radical is not
 2      on a direct sequela to visible injury.
 3           Runeckles and Vaartnou (1997; Figure 9-9c) discovered a signal by subtracting other EPR
 4      signals of the leaf, which seemed to be due to an O3 reaction with plant material, using 0.48 ppm
 5      O3. This  difference signal looked very much like O2 .  At a lower concentration, they observed
 6      that this signal still occurred but accumulated more slowly. Both bluegrass and ryegrass leaves
 7      seemed to saturate after about 5 h of exposure at 22 to 28 units of signal, while radish leaves
 8      reached a maximum of 7 units at 3 h and then declined. The problem, which is typical of any of
 9      these methods, was that the detached leaf had to be rolled and placed into the EPR detection
10      cavity. Reichenauer et al. (1998) also detected an undefined free radical signal that seemed to be
11      related to a Mn(II) spectrum.  The Nandu and Perlo cultivars  of wheat were more sensitive to O3
12      than Extradur (according to growth rate and closure of stomata under an O3 exposure of 80 ppb
13      for 8  h/day, 7 days/week over 100 days), and these more  sensitive cultivars had a greater, but
14      insignificant (P = 15%) EPR signal. Thus, data showing any  production of a free radical must be
15      approached with some skepticism.
16           With an O3 delivery rate of about 25 |iM/h (Table 9-6),  only 250 jiM would be found after
17      a full day, if all of the O3 were stable.  While the use of free radical traps is the best method
18      available  to observe any build-up of radicals, the traps  are not as specific to individual radicals.
19      Currently, studies should be looking for  hydroxyl radicals, superoxide, hyroperoxides, ethylene
20      radicals, and ascorbate radicals.
21
22      9.3.3.3 Products of Ozone
23           Ozone should reach a certain concentration in the substomatal cavity, which is dependent
24      upon its entry speed and its reactivity with the wall constituents.  Once near the apoplastic space,
25      O3 moves in two different pathways (Figure 9-6).  It  can react with constituents that are within
26      the wall as a gas in reaction 1A (path 1); or it can solubilize into  a water space and travel to
27      another region within the water space and react through reaction 2A (path 2).
28
29      Hydrogen Peroxide
30           Hydrogen peroxide, until recently was thought to be purely a toxic compound for cells.
31      However, it is now clear that it functions as a signaling molecule in plants and mediates

        January 2005                              9-48       DRAFT-DO NOT QUOTE OR CITE

-------
 1      responses to abiotic and biotic stressors (Figure 9-10). Generation of H2O2 is increased in
 2      response to various stressors, implicating it as key factor mediating the phenomena of
 3      acclimation and cross tolerance, in which exposure to one stressor can induce tolerance of
 4      subsequent exposure to the same or different stresses (Neill et al., 2002). The signaling response
 5      to attack by invading pathogens using H2O2 has been described (Mehdy, 1994; Simon-Plas et al.,
 6      1997).  The reactions leading to hypersensitive cell death are caused by a pathogen recognition
 7      step (Figure 9-10a), probably due to the plant cell wall releasing oligosaccharides in response to
 8      the pathogen enzymatically breaking down the cell wall to penetrate it.  A feed-forward step in
 9      which H2O2 increases the level of benzoic acid leads to the activation of the hydroxylase step in
10      the production of salicylic acid and to a feed-back step in which the salicylic acid increases the
11      production of H2O2 (Leon et al., 1995).
12           An elicitor, e.g., a bacterial or fungal pathogen, induces a cascade of reactions within a cell
13      (Figure 9-1 Ob). Some of the lipid reactions are thought to be due to the opening of the Ca2+
14      channels and the alkalination of the cell wall region.  The oxidative burst due to H2O2 production
15      is believed to lead to the transformation of a small population of lipids into jasmonic acid, which
16      is a secondary messenger.
17           Hydrogen peroxide also has an oxidative role in lignification (Schopfer, 1994). In the
18      interaction of lignification and the beginning processes of hypersensitivity, pectinase produced
19      by the pathogen disrupts pectin and dissolves the cell wall.  Fragments of the dissolved cell wall
20      trigger an increase in the transcription of peroxidases within the remaining cell wall,  leading to
21      lignification, which is a cross-linking of the  cell wall that does not use pectin.  This prevents
22      further pathogen disruption of the wall and reduces its further entry into the plant cell.
23           It is believed that  the first species generated through a one-electron reduction of molecular
24      oxygen is SO2 . That generation is carried out using a cytochrome b6 by the NAD(P)H oxidase
25      located on the cell membrane (Auh and Murphy, 1995).  In the acid region of the cell wall, SO2
26      is converted by a protonation and dismutation to H2O2.  The induced oxidative burst is believed
27      to play a role in stimulating the Cl"  and K+ efflux, generating an alkalinization of the
28      extracellular space  (Cazale et al., 1998). In the wall region, H2O2is not especially toxic, as no
29      necrosis was reported in tobacco when 500 mM peroxide was infiltrated into the leaf tissue.
30      However, the production of salicylic acid and benzoic 2-hydroxylase can be induced with only
31      30 and 0.3 mM H2O2 respectively, indicating some metabolic signaling  (Leon et al., 1995).

        January  2005                              9-49        DRAFT-DO NOT QUOTE OR CITE

-------
               a.
                                   Pathogen Recognition

                                              I
                                  Free Radical Generation
                             	H2O2 Generation

                                              I
                             Accumulation of Benzole Acid
                                        ^ Benzole Acid
                                 2-Hydroxylase Activation

                                              I
                            — Salicylic Acid Accumulation
                                              I
                    Systemic Salicylic Acid Increase
                                                Defense Activation
                                                       Hypersensitive Cell Death
                                                Local Resistance
                                                       Global Resistance
/ell Membrane C
Ascorbate
Peroxides
OH" \
1 *
\ / „ .(
G Proteins 	 	 Open Ca2+ 	 Phosphorylation * (
(Rac2/Rap1A) Channels Cascade %
I
Pathogen
/ Kill
/ t i
•( "Oxidative
+ Burst
Cell Wall
->. Structura
Changes
/ NADPH Hydroperoxides
H+ Oxidase ./
Jasmonic
/ Acid
                                    Influx
                                  Signal Transduction
                                                             Transcription
                                                                             Defense Genes
       Figure 9-10.  Pathogen-Induced Hypersensivity.  (a) The reactions leading to
                     hypersensitive cell death and the formation of a global response of salicylic
                     acid, (b) The cascade of the elicitor-induced reactions within the cell.
1

2
On the other hand, 1M H2O2 infiltrated into soybean will generate lipid peroxidation after 1 h

with a peroxidation rate of 15 nmol/g-FW h (Degousee et al., 1994). Cells react to the system5
               Soybean suspension cells were inoculated with Pseudomonas syringae pv syringae, which generate an
       active oxygen response. Light emission by luminol, reacting with H2O2, was the assay for the peroxide.
       January 2005
                                            9-50
DRAFT-DO NOT QUOTE OR CITE

-------
 1      and generate peroxide scavenging compounds within 1.5 to 2 hours, which appear to "mop up"
 2      the excess H2O2 (Baker et al., 1995).
 3           After O3 exposure in birch, H2O2 has been found in the wall (Pellinen et al., 1999).
 4      By using CeCl2 as a cellular stain for H2O2 (as a cerium perhydroxide precipitate), Liu et al.
 5      (1995) observed a gradual development of stain after 8 h of O3 exposure (at 150 ppb).  After 2 h
 6      exposure, H2O2 stain was visible on the surfaces of both sets of mesophyll cells. Accumulation
 7      of H2O2  stain continued for 16 h after exposure, suggesting a triggered-reaction rather than O3
 8      decomposition itself. H2O2 stain was present in the mitochondria, peroxisomes, and cytoplasm,
 9      but not in the chloroplast.  If methyl viologen (MV) was given to the leaves and then the leaves
10      were exposed to light, H2O2 stain could be observed within the chloroplast. This indicated that
11      the stain worked within the chloroplast if H2O2 were generated by the Mehler reaction
12      (MV+O2 ).  Thus, apparently, for birch, O3 exposure does not generate excess H2O2 within the
13      chloroplast.  Furthermore, these sets of experiments indicate that O3per se does not generate the
14      H2O2, but rather triggers stress-related H2O2 formation similar to what occurs in a pathogen
15      attack (the Reactive Oxidative Species or ROS reaction).
16           The presence of higher than normal levels of H2O2 within the  apoplastic space is a potential
17      trigger for the normal, well-studied pathogen defense pathway.  Figure 9-1 Ob depicts such a
18      pathway and suggests that all the events and activation of pathways/genes caused by pathogen
19      defense could be observed when plants are fumigated with O3.  These events in Figure 9-1 Ob will
20      be alluded to in later sections.
21           H2O2 has been linked to the hormone ABA-induced closure of the stomata by activating the
22      calcium  influx  in guard cells (Pell et al., 2000). The addition of H2O2 at a level of only 5 mM to
23      a guard cell preparation will cause a dramatic increase (ca. 9*) in electrical current at the
24      hyperpolarizing potential of-200 mV. Amounts as low 50 mM H2O2 will cause a less but still
25      sizable increase. Membrane stability is unaffected by the H2O2 and the activation of the channel
26      requires  only about 2 to 3 minutes. Pell et al. (2000) also found that ABA induced the
27      production of H2O2through ROS accumulation (also see Zhang et al., 2001).
28           Certain levels of ABA within the leaf lead to stomatal closure. The inactivation of a
29      phospho-tyrosine-specific protein phosphatase (ABI2) is an inhibitor of stomatal opening
30      induced  by ABA but that enzyme is inhibited by H2O2 (Meinhard et al., 2002). This means that
31      H2O2 shifts the sensitivity of the stomatal opening to ABA (Figure 9-11), making the stomatal

        January 2005                              9-51        DRAFT-DO NOT QUOTE OR CITE

-------
                                             ABA
                                                                    Level
                                                           9s
                                                                          Control
                                           Stomatal
                                            Closure
                                                                    [ABA]
       Figure 9-11.  The interaction of H2O2 and Ca2+ movements with ABA-induced stomatal
                    closure.  It is well known that certain levels of ABA within the leaf lead to
                    closure of the stomata within the leaf.  That level, however, can be shifted to
                    make closure more or less sensitive to a given level of ABA.  Recently it has
                    been shown that H2O2 (externally or produced by the plant) within the cell
                    wall region can shift that sensitivity. Here ABA stimulates the production of
                    H2O2, which in turn increases the rate of Ca2+ moving from the wall region
                    into the cytoplasm. That shift in internal Ca2+ level increases the closure of
                    the stomata. Hydrogen peroxide also blocks the activation of a polypeptide
                    (ABI2) that inhibits stomatal closure seemingly induced by ABA.
       Source: From Assmann (2003); Assmannand Wang (2001); and Zhang et al. (2001).
 1     complex more sensitive to ABA. Thus, for a given level of ABA present in the guard cell
 2     complex due to environmental factors (e.g., low humidity, high air temperature, or low soil water
 3     potential), the generation of H2O2 would (by inhibiting ABI2) induce a closure of the stomata by
 4     increasing the sensitivity of the guard cells to ABA. In the past, it has been difficult to
 5     understand why O3 would often decrease conductance in some cases, but not always (Heath,
 6     1994b). This interaction between H2O2 and ABA could help understand this complexity.
 7
 8     Ethylene Reactivity
 9          Ethylene (ET) is produced when plants are subjected to biotic stressors e.g., attacks by
10     insects, fungi, and bacteria or abiotic stressors such as wounding or environmental stressors such
11     as heat, cold, or oxidative stress and O3. If an O3 stress has induced a wounding response with
12     ET release, then ET within the substomatal cavity could react with O3, generating some
13     relatively noxious chemicals (see Figure 9-6). The relationship between O3 injury and wounding
       January 2005
9-52
DRAFT-DO NOT QUOTE OR CITE

-------
 1      is supported by the observation by Mehlhorn et al. (1991) that an inhibitor of ET formation,
 2      AVG (an inhibitor of ACC synthase, a committed step to ET production), would block ET
 3      formation and inhibit visible injury.  Other studies with polyamine (which is closely linked to
 4      ET production), including those of Ormrod and Beckerson (1986) who fed polyamines to the
 5      transpirational stream and prevented visible injury, suggested a close involvement of both
 6      pathways to the production of visible injury. Both the lack of ET production or an increased
 7      level of polyamines slowed or prevented visible injury.
 8          This concept was taken another step by Langerbartles (Langebartels  et al., 1991). The
 9      linkage to the pathogen wound responses and visible injury is well established (Sandermann,
10      1996). Sandermann (1998) used a system of Bell B and W3 tobacco, plants with differential O3
11      sensitivities, in which the O3 exposure  level was  chosen such that the sensitive cultivar was
12      injured, while the tolerant one was not. This led  to a marvelous control which could be used to
13      their advantage. They followed a time sequence  to show that the rise of varied systems followed
14      the same order as for a pathogen attack (Heath, 1994a).
15          More recent studies, however, indicate that O3 responses resemble components of the
16      hypersensitive response (HR) observed in incompatible plant-pathogen interactions (Sanderman
17      et al., 1998).  The similarity to the HR  response may be related to the occurrence of ROS in the
18      apoplast. The O3-derived ROS apparently trigger an oxidative burst in the affected cells by an as
19      yet unknown mechanism.  An oxidative burst is similar to one of the earliest responses of plants
20      to microbial pathogens and is an integral component in HR-related cell death (Overmyer et al.,
21      2000).
22          In plants exposed to  O3, ET synthesis is a result of the specific ET induction of the genes
23      encoding 1-aminocyclopropane-l-carboxylase synthase (ACS), one of the fastest and most
24      obvious responses to O3, which has been mechanistically linked to the regulation of O3 lesion
25      formation.  Biosynthesis of ET inhibited with ACS inhibitors significantly reduced the induction
26      of lesion formation in plant leaves exposed to O3 concentration (Vahala et al., 2003; Melhorn
27      and Welburn, 1987; Mehlhorn et al., 1991).  Ethylene biosynthesis correlates best with O3
28      (Vahala et al., 2003; Overmyer,  2000). These data support the concept that elimination of ET
29      formation will prevent visible injury.
30
        January 2005                              9-53        DRAFT-DO NOT QUOTE OR CITE

-------
 1      Ethylene-Interaction with Injury and Conductance
 2           Increased ET production by plants exposed to O3 stress was identified as a consistent
 3      marker for O3 exposure decades ago (Tingey et al., 1976). They exposed more than 20 plant
 4      species and cultivars to O3 to determine whether the production of O3-induced stress-ET could be
 5      used to determine differences in plant sensitivity to O3.  Their studies suggested that increased
 6      production of stress-ET correlated well with the degree of foliar injury that developed within
 7      hours or days after O3 exposure.  The amount of ET released was exponentially related to the O3
 8      exposure. Furthermore, the amount of O3-induced ET declined with repeated exposure,
 9      indicating an acclimatization to O3.  This acclimatization effect associated with repeated
10      wounding has not yet been well described.  The release of wound-induced ET is not linear with
11      time, but declines after the initial response (Stan et al.,  1981), as is also seen after O3 exposure
12      (Stan and Schicker, 1982). The stress-induced ET production correlates better with O3 exposure
13      level than with exposure duration. In other words, peaks of high O3 (rather than accumulated
14      dose) generate a higher rate of ET release, at least for a single O3 exposure under an acute dose.
15           The production of ET after an  O3 exposure is thought to be a typical wounding response
16      (Tingey et al., 1975).  Prevention of ET release may prevent the formation of visible injury
17      (Mehlhorn and Wellburn,  1987). However, the question arises as to whether this effect was
18      limited to the prevention of visible injury or if the chemicals used to prevent ET release closed
19      the stomata.  Using Glycine max L., Taylor et al. (1988b) showed clearly that AVG did not
20      necessarily close stomata nor inhibit carbon assimilation/>er se.
21           The correlation of ET release with O3-induced visible injury was likewise shown in pea
22      cultivars (Dijak and Ormrod, 1982). With O3 exposure (generally 6 h at 0.3 ppm), the stomata
23      closed by -50% within 3 hours after a dose of 3 x  10~5 mol cm"2 (with an average rate of
24      2 x io~9 moles cm"2 s"1, as calculated from their data).  Both sensitive and insensitive cultivars
25      had a visible-linked-injury ET release, but sensitive cultivars scored higher both in visible injury
26      and in ET release after a given exposure.
27           Gunderson and Taylor (1988, 1991) used exogenous ET to alter the gas exchange of
28      Glycine max and found an exponential, but not simultaneous, decline of both stomatal
29      conductance and carbon assimilation with ET.  Interestingly, the exogenous ET caused a slight
30      rise in difference of CO2 within and  without the leaf, indicating a lowering of internal CO2,
31      which was not observed in the experiments of Farage et al. (1991) for O3 exposure. Ethylene

        January 2005                             9-54        DRAFT-DO NOT QUOTE OR CITE

-------
 1      does inhibit both stomatal conductance and carbon assimilation to some extent (Taylor et al.,
 2      1988b).  Thus, one could postulate that O3 generates a wounding response with a production of
 3      ET, which would, in turn, generate the change in stomatal conductance and photosynthesis.
 4      Clearly, these multiple events may have confounded some earlier studies.
 5
 6      9.3.3.4  Antioxidants Within the Apoplastic Space
 7           The first line of defense against O3 is a closure of the stomata to exclude its uptake. This is
 8      counter-productive for efficient photosynthesis, but some amount of closure limits the rate of O3
 9      deposition into the leaf tissue to allow for a secondary line of defense to detoxify the O3. The
10      secondary line of defense involves a range of antioxidants, which are highly reactive to the types
11      of chemicals that can be generated by O3.  Several antioxidant proteins are stimulated by O3
12      in Plumbagini folia, including glutathione peroxidase (GSH-PX), SOD, and catalase. The
13      timescales for changes in their levels vary:  some rise rapidly, while others rise more slowly.
14      The pattern of changes in these particular proteins varies greatly among different species and
15      conditions.
16
17      Ascorbate Within the Cell Wall
18           Most of the recent reports indicate that ascorbate within the cell wall is the real first line of
19      all defense. Ascorbate within the wall declines when the tissue is exposed to  O3 (Luwe et al.,
20      1993; Moldau, 1998; Turcsanyi et al., 2000; Zheng et al., 2000). This decline appears to be
21      closely linked to the amount of O3 penetrating the leaf tissue.
22           It has long been suspected that intracellular antioxidants play a role in preventing O3-
23      induced injury to plant cells.  Variation in the types of biochemical compounds present in the
24      apoplastic space can give rise to a multiplicity of reactions with O3, but the predominant
25      biochemical species is ascorbate.  Ascorbate is water-soluble, present in the solution where O3
26      can dissolve, and is highly reactive. Unfortunately, a variety of antioxidants are found
27      throughout the cell and any measurements of one particular type within the total leaf tissue can
28      give misleading results. For example, ascorbate is present within the cell wall, within the
29      cytoplasm (Moldau, 1998; Burkey,  1999), and within the chloroplasts (Law et al., 1999); and
30      ascorbate can move between  the cytoplasm and the cell wall with relative  ease (Figure 9-12;
31      Bichelle et al, 2000).  The total of all ascorbate pools are measured when the tissue is ground and

        January 2005                              9-55        DRAFT-DO NOT QUOTE OR CITE

-------
                                         Cell
                       Apoplasm   Membrane    Cytoplasm
                   a.
                                                Ma late
                                                         NAD
                     NADH
                               > De hydro-.
                                ascorbate
                                Ascorbate
                                 HI  Cytb
                                Dehydro-
                                ascorbate
          -^/AscoTbateJ


II trai sporter(s)         \(

          	De hydro-  A
            ascorbate
                                              Dehydro-
                                              ascorbate
                                                                  [ Ascorbate
                                                                   Dehydro-
                                                                   ascorbate
                    b.
                            Ascorbate
                           Monodehydro-
                             ascorbate
                               Ascorbate
                          Carrier
Dehydro-
ascorbate
                                                Ascorbate

                                                      Cytochrome b
           Monodehydro-
             ascorbate
                                                  NADH
                                                      Oxidoreductase
                                                  NAD+
                                              Ascorbate
                                              De hydro-
                                              ascorbate
                                                           GSH
                      •GSSG
                                                                     NADPH
           NADP+
                                                    Dehydro-   Glutathione
                                                    ascorbate   Reductase
                                                    Reductase
Figure 9-12.  The reaction of ascorbate within the apoplasm of the cell wall and its
              ultimate reduction/oxidations.
              (a)   Movements of reducing power (from Dietz, 1997).
              (b)   The use of glutathione to maintain the level of ascorbate
                    within the cell wall region (from Horemans et al., 2000).
January 2005
         9-56
DRAFT-DO NOT QUOTE OR CITE

-------
 1      assayed. If the cell wall ascorbate drops by 50% due to O3 exposure but all other tissue
 2      concentrations remain the same, the measurement of the total loss is dependent upon the amount
 3      of ascorbate within the cell wall.  Turcsanyi et al. (2000) showed that, compared to the
 4      concentration of apoplastic ascorbate, the rest of the cells contained about 38 times as much.
 5      So a 50% loss of apoplastic ascorbate would be converted into only 2 to 3% loss of the total
 6      ascorbate.
 7           The ascorbate deficient Arabidopsis thaliana mutant has proven to be a powerful tool in
 8      furthering the understanding of ascorbate biosynthesis in plants (Smirnoff et al., 2001).  Three
 9      classes of mutants were formed when A thaliana seed was mutagenized with ethyl
10      methanesulfonate:  (a) those deficient in SOD, (b) those that failed to accumulate more
11      antioxidant proteins upon increased O3 exposure, and (c) those that were deficient (but not
12      depleted) in ascorbate. The low-ascorbate mutant type had 50 to 60% less ascorbate than  the
13      wild type and displayed more foliar injury.  This mutant is involved with the coding of the
14      GDP-D-Mannose pyrophosphorylase enzyme6 in the Smirnoff-Wheeler pathway for ascorbate
15      biosynthesis.  Smirnoff et al. (2001) also suggested that other pathways can produce ascorbate
16      without relaying upon the pyrophosphorylase step, but most probably at a slower through-put
17      rate, because any fully ascorbate-deficient mutant would be lethal, perhaps because of ascorbate
18      use as a cofactor rather than its antioxidant properties.
19           The ascorbate peroxidase (APX, which uses ascorbate to detoxify peroxides) family
20      consists of at least five different isoforms, with isozymes in the apoplastic and cytosolic space.
21      Furthermore, most forms of ascorbate can move through the plasma membrane (Figure 9-12;
22      Bichele et al., 2000), making the levels of all forms of ascorbate interdependent and able to at
23      least partially influence each other.  Dehydroascorbate (DHA) can be broken down into other
24      smaller fragments easily in vivo and represents a continuous loss of ascorbate from varied parts
25      of the cell if ascorbate is allowed to remain in the oxidized form in some regions.  In fact,  the
26      turnover rate in leaves is estimated to be from 2 to 13% / hour, depending upon species and
27      developmental age (Smirnoff et al., 2001). There are apparently three pathways for ascorbate
              6 EC 2.7.713, Mannose-1 phosphate guanylyltransferase; mannose + GTP —> GDP-mannose + ppi;
        this product leads into cell wall polysaccharide synthesis and protein glycosylation through GDP-galactose and
        GDP-fucose and ultimately, through Galactose, into ascorbate synthesis.

        January 2005                              9-57        DRAFT-DO NOT QUOTE OR CITE

-------
 1      turnover (Figure 9-12a).  The typical reaction is a reduction of DHA into ascorbate from which
 2      an oxidative step generates DHA. Pathway I requires a reductive step using NADH external to
 3      the plasma membrane generated from internal malate using a malate/oxaloacetate transporter.
 4      Pathway II uses a direct transporter of ascrobate/DHA.  Pathway III moves the electron(s)
 5      required through a cytochrome b system, maintaining two separate pools of ascorbate/DHA
 6      within the cytoplasm and within the wall. Each of the pathways (Dietz, 1997) represented by
 7      Roman Numerals in the Figure 9-12, require only one NAD(P)H molecule to reduce the DHA
 8      molecule back to ascorbate.  However, the transport properties and redox potential of the cell
 9      differ for each pathway. The efficiency of the reduction of DHA is dependent upon the redox
10      coupling and the region in which the chemical species is located.
11           Turcsanyi et al. (2000) exposed broad bean (Viciafabd) grown under two regimes in
12      duplicate controlled chambers: charcoal/Purafil filtered air (CFA) or (CFA) plus 0.075 ppm O3
13      for 7 h/day for 28 days (chronic exposure) or exposed to 0.150 ppm for 8 h (acute exposure).
14      Responses of the two set of plants were similar except for stomatal conductance,  which was 50%
15      lower in the chronic exposure plants.  Plants grown under acute exposures developed visible
16      injury, while plants grown under chronic conditions developed no visible injury.  Within an hour
17      of the start of the acute exposure, the stomatal conductance was reduced by nearly 40% and
18      assimilation was reduced by nearly 18% in the clean air plants; a reduction in conductance was
19      only 21% and assimilation 16% in the plants subjected to chronic O3 exposures.  The
20      assimilation was affected similarly in both cases, while  the conductance showed less of a
21      percentage drop in the chronic O3-exposed plants, beginning at a lower O3 level.  The similarity
22      of the assimilation indicated that the stomata were not limiting assimilation in either case before
23      acute exposure.  More to the point, the decline in ascorbate in the apoplastic space due to the O3
24      exposure was ".. .more often than not, on the borderlines of statistical  significance."  However, a
25      30% decline in ascorbate after 4 hours of acute O3 in both cases was observed.  This lack of
26      significance may be due to a relatively large standard error of the data, which in turn may be due
27      to the difficulty of extracting and measuring ascorbate from the apoplastic space in quantitative
28      terms.
        January 2005                              9-58        DRAFT-DO NOT QUOTE OR CITE

-------
 1           The chemical reaction7 of ascorbate and O3 is given by the molecular rate constant of
 2      4.8 x 1071VT1 s"1.  This is some 50,000x the rate constant for ET. Of course, it depends upon the
 3      relative concentration of ascorbate and ET, but it is likely that ascorbate is in higher
 4      concentration than ET.  One would then expect that the rate reaction of ascorbate with O3 would
 5      greatly dominate any possible reaction of O3 with ET. For a concentration of ascorbate in the
 6      range of 1  mM and for an O3 concentration of about 0.1 ppm or 4.2 x 1CT9 M, the detoxification
 7      rate would be 4.8  x 107 x i(r3 x 4.2 x l(T9 M s"1 = 2.0 x l(T4 M s"1. Turcsanyi et al. ( 2000)
 8      calculated an O3 flux of about 1.6 x io~9 moles m"2 sec"1.  With a wall thickness of 0.12 x  io~6 m
 9      and all the O3 flux going into the wall region, this would give about 1.3 x 10~2 mol m"3 s"1 or
10      1.3 x 1Q~5  M s"1 flux, which is less than 10% the detoxification rate.
11
12      Glutathione
13           Many of the initial studies of O3 exposure used high concentrations and measured only the
14      total sulfhydryl  contents of the tissues. For example, in some of the earlier work, exposures of
15      tobacco to 1 ppm O3 for 30 min induced a 15% loss of the total sulfhydryls (0.74  |imole/g-FW;
16      Tomlinson and Rich, 1968).  These results are similar to other studies at high O3 levels (Dugger
17      and Ting,  1970). It is now suspected that severe injury in their studies resulted in a massive
18      collapse of the cells releasing most of their internal constituents.  Much of the oxidation thus
19      observed may have been the result of chemical oxidations of the O3 that subsequently entered the
20      damaged tissue. Even under milder conditions, changes in sulfhydryl components have still
21      been noted and any sulfhydryl on the surface of the cell would be at risk due to its high reactivity
22      with O3 (Mudd et al., 1969; 1997a). For example, the level of sulfhydryl  compounds within the
23      protein of isolated chloroplasts declined about 66% when the chloroplasts were subjected to O3
24      (about  1 jimole  O3) exposure (Mudd et al., 1971).
25           At this stage, it is important to note that there are inherent problems with metabolic studies
26      of full tissues. The first is that most organs have several different types of tissues. For example,
27      leaves have, at the minimum, epidermal and vascular tissues and two types of mesophyll cells.
28      Each type  of cell may be metabolizing quite differently, producing very different levels of
               7These chemical rate constants are those constants within a bulk solution. In the apoplasm, the possibility
        exists for the chemicals to be preferentially oriented near a surface; so the constants may not be the same as for
        bulk solutions.

        January 2005                              9-59        DRAFT-DO NOT QUOTE OR CITE

-------
 1      metabolites and enzymes. Furthermore, most pathways are well regulated and after any small
 2      disruption, the pathway tends to return to near its former stability.  Changes in the level of
 3      enzymes are likewise difficult to measure. Many enzymes function below their maximum
 4      activities. Their speeds of reactions are often increased through regulation, rather than through
 5      the production of more enzyme.
 6           Glutathione is a three-amino acid peptide, which has antioxidant properties due to its free
 7      reducing sulfhydryl group (G-SH). Glutathione is generally kept in its reduced form by
 8      glutathione reductase (GR) with the reaction:
 9
                             GS - SG(oxidized) + 2e~ + 2H+ -> 2GSH                      (9-2)

10
11      GR has six isoforms8 within the chloroplast and six isoforms outside.  The optimum activity
12      occurs at pH 7.8, suggesting it is located within the stroma of the chloroplast or the cytoplasm
13      rather than in the cell wall, which is at pH 4-5  (Madamanchi et al. (1992). Clearly, an increased
14      expression of GR (generated through transgenic implants) is important within the chloroplast to
15      prevent of some oxidations9 (Aono et  al., 1995; Foyer et al., 1995).
16
17      Catalase
18           Catalase, even though it breaks down H2O2, does not appear to protect plants from O3
19      exposures. Two principle reasons may cause this lack of reactivity: (1) catalase has a high Km
20      for H2O2 and a low rate coefficient and (2) seems not to occur within the cell wall regions but
21      rather in the cytoplasm and peroxisomes (Buchanan et al., 2000).  Only a few reports suggest
22      that catalase is increased by exposure  to O3 (Azevedo et al.,  1998).  Unfortunately H2O2 induced
23      by some forms of wounding in mesophyll cells can lead to induction of an increase in GSH and
24      the transient production of catalase (Vanacker et al., 2000).  In general, it seems that catalase is
                An isoform is the same enzyme, with the same structure and perhaps within the same organelle, but its
        promoter region has different DNA codes.  Thus, each protein segment is induced by different signals, and so its
        enzyme can be formed in response to different environments.  This is in contrast to isozymes, which classically are
        similarly  reacting, but structurally different, enzymes in different compartments.
               9Typically this protection is observed in the paraquat sensitivity of plants. In this assay, added paraquat,
        the herbicide which intercepts electrons from the reducing end of photosystem I in the chloroplast, caused
        oxidations, chlorophyll loss, and death due to the buildup of superoxide and peroxides.

        January 2005                               9-60        DRAFT-DO NOT QUOTE OR CITE

-------
 1      not really involved primarily in the defense of the cell due to O3 attack but rather may be a
 2      secondary response.  The reaction of catalase (Scandalios, 1993) is as follows:
 3
 4                      2H2O2 -> 2H2O + O2      K = 1.7 x ICT7 N/T1 sec'1                 (9 3)

 6
 7      Superoxide Dismutase
 8           The varied compounds that O3 can produce upon entering an aqueous solution are very
 9      similar to those involved in the HR when plants are infected by an avirulent pathogen (Figure
10      9-10). The sequence of the plant response to the pathogen is: (1) recognition of the gene
11      products of the pathogen by the plant (elicitor), (2) generation of an immediate phytoresponse to
12      attempt to localize the attack and its products, and (3) generation of a systemic acquired
13      resistance (SAR) to subsequent attack by the pathogen. Inducible defense responses are
14      phytoalexin synthesis and production of pathogenesis-related proteins (PR). One aspect of this
15      total response is the production of O2 and H2O2 by the cell (Lamb and Dixon, 1997). The
16      elicitor can generate a transient alkalinization of the apoplast, up to pH 7.2, caused by a lowering
17      of the H+-pump rate and a increase in the H+-influx/K+-efflux exchange.  Other effects include:
18      a weak accumulation of transcripts for PAL (phenylalanine lyase); a larger and rapid induction
19      of glutathione S-transferase, GSH-PX; oxidative cross linking of cell wall proteins which is
20      blocked by ascorbate acid; generation of localized apoptosis; and rapid influx of Ca2+, which
21      activates apoptosis among other pathways (Lamb and Dixon, 1997). These effects seem to be
22      very similar to those induced by O3 exposure (Sandermann, 1996; 1998).
23           The putative antioxidant enzyme SOD (Equation 9-4 and Table 9-8) catalyzes the
24      oxidoreductase reaction, which eliminates SO2 by dismutation (Bowler et al., 1992):
25
26               2O2- + 2H+ —SOD  > H2O2 + O2      K = 2.4 x 109 M'1 seer1.           (9-4)
27
28
29           The number, as well  as the activity, of isozymes of each type of SOD in Table 9-3 can vary
30      with plant species. However, the isozymes that have been tabulated are Cu-Zn SODs, in cytosol
31      and chloroplast;  Fe-SOD, active in chloroplast stroma, and Mn-SOD, active in mitochondrial

        January 2005                              9-61         DRAFT-DO NOT QUOTE OR CITE

-------
                        Table 9-8.  Superoxide Dismutase Isozymes and Isoforms
Reaction: 2H+H
Isozymes
Cu-Zn


Fe


Mn
2 O2 - H2O2 + O2
M.W.
20kDa


23kDa


23kDa

Isoforms
csdl
csd2
csd3
fsdl
sd2
fsd3
msdl

Cytolocation

Plastid
Peroxisomal
Mitochondrial

Plastid
Mitochondrial
 1     matrix (Karpinski et al., 1993). In the experiment demonstrating the activation of varied SODs,
 2     there were three Cu-Zn SOD (csdl, csd2, csd3\ three Fe-SODs (fsdl,fsd2,fsd3), and one
 3     Mn-SOD (msdl) (Kliebenstein et al., 1998).  Ozone sensitivity was determined by exposure of
 4     plants to 8 h of 0.33 ppm of O3. csdl induced by O3 and UV-B was one of the earliest SOD
 5     increases and most pronounced responses for mRNA and protein.  Also, some increase in csd3
 6     (thought to be peroxisomal) was induced when the plants were exposed to a high-intensity light
 7     pulse; msdl was unresponsive to the environmental stressors used here, including O3; and csd2
 8     (thought to be chloroplast) showed little increase. The fsdl isozyme (present in the apoplasm)
 9     showed a slight decrease.  On the other hand, an early report on snap beans in which the
10     experimenters used EDU, N-[2-(2-oxo-l-imidazolidinyl)ethyl]-N-phenylurea (Carnahan et al.,
11     1978; Kostka-Rick and Manning, 1993) to prevent visible injury by O3, 4 h O3 exposure at
12     0.45 ppm was correlated with an increase in general enzyme activity of SOD, i.e., the level rose
13     nearly 2.5x in 2 weeks at a level of 50 mg EDU per pot (Lee and Bennett, 1982). It is believed
14     that EDU may induce SOD, which then protects the plant.  While gross assays of enzyme
15     activity have not proven to be very useful in understanding the mechanism of O3 action, in a
16     well-crafted, long-term study involving ponderosa pine clones. Benes et al. (1995) stated that
17     "changes in antioxidant enzyme activity were not a consistent response to the O3 fumigation, but
18     when observed, they occurred  most often in the O3-sensitive clone  and  in symptomatic,

       January 2005                             9-62        DRAFT-DO NOT QUOTE OR CITE

-------
 1      fumigated branches.. .total (intra- and extracellular) activities of the antioxidant enzymes did not
 2      appear to be good indicators of O3 tolerance...."
 3           Ozone exposure (70 ppb for 7 h/day for 14 to 42 days of exposure)10 caused an increase in
 4      POD and a decline in SOD with no change in APX. No GSH was detected, but the
 5      concentration of (ASC + DHA) was at 20 to 25  nmol/g-FW of extracellular fluid, compared with
 6      2.4 to 3.0 mmol/g-FW of cell fluid. Glutathione within the cell was only 100 to 170 nmol/g-FW
 7      of cell fluid. While these results are what one might expect for POD, the decline in SOD and
 8      lack of change in APX are not what would be expected if protection was provided by SOD and
 9      ascorbate. Yet as noted, because the rate of SOD reaction is many times higher than the rate of
10      O3 entry, there may be no pressure to increase the SOD level.
11           Some protection against visible injury (induced by 59 ppb daily mean O3 for 14 h/day for
12      7 days) was observed in genetically modified tobacco plants with excess chloroplast SOD11 (2 to
13      4 times higher), but less protection was observed in plants that had  an excess of mitochondrial
14      SOD (8* higher) (Van Camp et al., 1994). In all lines, the conductance of the leaves dropped
15      about 50%, compared with the unmodified plants.  There was a correlation with age of leaf (less
16      injury in younger leaves) that corresponded to that found in spruce  trees in which the amount of
17      SOD declined in relation to the longer that needles were held on the tree (Polle et al., 1989).
18      A slightly different study, however, found no O3 protection with varied SOD within the needles
19      (Polle and Rennenberg, 1991).  Interestingly,  in maize, the synthesis of SOD (any form) was not
20      stimulated by O3 exposure (at 0.50, or 0.75 ppm for 8 h, variable times thereafter) but was by
21      exposure to 90% O3 (Matters and Scandalios, 1987). It may be that this high level of O3 does not
22      affect the SOD, or perhaps it stimulates and degrades the enzyme simultaneously.
23           The conclusions to be drawn from these results are not obvious. There seems to be SOD (a
24      Cu-Zn form) present in the apoplastic space of some plants, but it does not necessarily rise with
25      O3 exposure.  Thus, either its concentration is sufficient to provide protection or it is not needed.
26      Over-expression of any SOD in other organelles may play a role, especially in the chloroplast
                At a level of 70 ppb, the concentration of O3 in air was about 3.06 x KT6 mol/m3, which with the
        conductance of 0.042 mol/m2 s, gives a flux rate of O3 of 1.27 x 10~8mole/m2 s. Converting the SOD rate of
        23 units/g-FW into a SOD rate within the apoplastic space of 6.9 x 10~3 mol/m2 s, or about 500,00 times the
        entry rate of O3.
                The SOD enzymes were from Nicotiana plumbaginifolia with appropriate transit sequence for targeting
        the correct organelle and expressed under control of cauliflower mosaic virus 35S promoter.

        January 2005                              9-63        DRAFT-DO NOT QUOTE OR CITE

-------
 1      (Cu-Zn or Fe forms); but it may be playing a secondary role due to other effects of O3 that
 2      generate conditions in which light can overload the chloroplast and generate detrimental
 3      circumstances, including the production of SO2 .  In addition, SOD is developmentally expressed
 4      in varied concentrations, so that long term exposure to O3 may alter each leafs developmental
 5      age and, in turn, alter what level of SOD is observed. In any case, SOD does not seem to be the
 6      primary antioxidant system to protect against O3.
 7
 8      Changes to the Plasmalemma
 9           Reports of "peroxidation" generally occur within unicellular organisms subjected to very
10      high levels of O3 (in Chlorella [Frederick and Heath, 1970)] and in Euglena [Chevrier et al.,
11      1990)]). Heath (1987) determined that by the time biochemical events were altered and MDA
12      was produced in Chlorella, little permeability remained in the cells and most metabolic
13      pathways were greatly disrupted by the subsequent loss of substrates. In fact, MDA production
14      was concurrent with a high O3 uptake by the cell,  indicating a complete opening of the cell and
15      associated with the concurrent inability to plate the cells on a glucose median (indicative of cell
16      death). Heath reached the conclusion that no one had proven that lipid oxidation was in any way
17      a part of the initial reactions of O3 with the cell, a conclusion confirmed by Mudd et al. (1997a).
18      An excellent review regarding the initial action induced by O3 within a plant (Kangasjarvi et al.,
19      1994) should be consulted. There is little data to  show that lipids are attacked by O3 in any
20      living system that was not previously severely injured by O3. Most of the  data suggesting lipid
21      attack by O3 has been demonstrated in plants subjected to O3 concentrations of 0.5 to 1.0 ppm for
22      several hours, during which gross wilting of the plant tissues usually occurs, suggesting extreme
23      water loss. It is not surprising that lipid and protein injury is observed under these conditions.
24      While those reports were useful in the 1960s and  1970s,  they are not especially insightful now
25      when ambient levels of O3 are rarely above 0.2 ppm.
26
27      9.3.4  Wounding and Pathogen Attack
28           The decline of an enzyme is more difficult to measure than the rise of a new enzyme; an
29      increase from 0 to 2% may be within the precision of any assay, but a decrease from 100% to
30      98% is often masked by simple variation of the assay. Thus, measuring enzymes, which are in
31      great abundance in pre-fumigated tissue, is a risky operation.  On the other hand, if O3 induces a

        January 2005                              9-64        DRAFT-DO NOT QUOTE OR CITE

-------
 1      general physiological change that has characteristics similar to other well-studied, stress-induced
 2      changes, then O3 studies could "piggy-back" onto those studies to gain insight into the full scope
 3      of metabolic alterations.  It is now becoming clear that wounding and pathogen attack of plants
 4      are similar to O3-induced changes in plants, and a reasonable hypothesis is that O3 must induce
 5      one or more of the first steps seen in the wounding/pathogen-attack response.
 6           Systemic acquired resistance (SAR) has been heavily investigated, and DNA probes have
 7      existed for some time for a series of expressed genes (see Table 9-9). Several enzyme classes
 8      are associated with O3 injury, including glucanases and peroxidases and others, such as the PR
 9      proteins and  chitinases.  Thus, strong evidence exists from enzyme function and genetic material
10      that O3 induces an activation of a SAR-like response.
11           Mehdy (1994) described a model of how an elicitor produced by the pathogen attack
12      activates a G-protein, which opens the inward-flowing Ca2+ channel.  The flow of Ca2+ into the
13      cytoplasm raises the internal level (at the jiM level) and activates a protein kinase that increases
14      the activity of the plasma membrane NAD(P)H oxidase and generates O2~. Superoxide
15      dismutase converts O2 into H2O2. Both O2 and H2O2 are responsible for the active oxygen
16      species response, which is believed to be a defense mechanism to kill the pathogen. In this
17      normal defensive reaction, a subsequent system induces either localized lipid peroxidation per se
18      or a membrane lipase to produce j asmonic acid or inositol triphosphate, which act as secondary
19      messages to activate the defense gene products.
20           Booker et  al. (2004) found that G-proteins might be involved in the perception of O3 in the
21      extracellular region using A. thaliana G-protein null mutants. The activation of a passive inward
22      flow  of Ca2+, e.g., by an O3-induced response, would serve the same function as activation of the
23      G-protein. Once the level of cytoplasmic Ca2+ rises, all else follows.  It is suspected that
24      exposure  of plants to O3 does just that, as Castillo and Heath (1990) demonstrated — the in vivo
25      fumigation of bean plants both inhibits the outward-directed ATP-requiring Ca2+ pump and
26      increases  the passive permeability of Ca2+. It was thought that the calcium transporter system
27      has a sensitive sulfhydryl group which, if oxidized, would alter normal Ca2+ movements.  In
28      addition, Dominy and Heath (1985) observed that the K+-activated ATPase (believed to be
29      involved in K+transport) was inactivated by in vivo exposure to O3 and that inactivation was
30      traced to a sensitivity sulfhydryl.  Mudd et al. (1996) argued that several amino acids are very
31      sensitive to O3, including any with an exposed sulfhydryl. Thus, the O3-induced change in Ca2+

        January 2005                             9-65         DRAFT-DO NOT QUOTE OR CITE

-------
         Table 9-9. Gene Families and cDNA Clones Used as Probes for SAR (Ward et al., 1991)
        Probe
                   Relevant Properties of Encoded Protein
        Reference
        PR-1



        PR-2


        PR-3


        PR-4


        PR-5



        PR-1 basic

        Basic class III
         chitinase

        Acidic class III
         chitinase

        PR-O'


        Basic, glucanase


        Basic chitinase

        SAR 8.2
                   Acidic, extracellular; function unknown most
                   abundant PR protein in tobacco; >90% identical to
                   PR-lbandPR-lc

                   Acidic, extracellular b-1,3-glucanase, >90% identical
                   to PR-N and PR-O

                   Acidic, extracellular chitinase; also known as PR-O;
                   >90% identical to PR-P

                   Acidic, extracellular; unknown function; homologous
                   to C-terminal domain of Winl and Win2 of potato

                   Acidic, extracellular; homologous to thaumatin and
                   bifunctional amylase/proteinase inhibitor of maize;
                   also known as PR-R or PR-S

                   Basic isoform of acidic PR-1

                   Homologous to cucumber chitinase (Metraux et al.,
                   1989), structurally unrelated to PR-3

                   Extracellular; approximately 60% identical to basic
                   isoform

                   Acidic, extracellularb-l,3-glucanase; approximately
                   55% identical to PR-2 group

                   Vacuolar; approximately 55% identical to PR-2 group
                   and PR-O'
        Payne etal. (1988b)



        Ward etal. (1991)


        Payne etal. (1990a)


        Friedrichetal. (1991)


        Payne etal. (1988a)



        Payne etal. (1989)

        Lawton et al. (manuscript in
        preparation)

        Lawton et al. (manuscript in
        preparation)

        Payne etal. (1990b)


        Shinshietal. (1988)
                   Vacuolar; approximately 65% identical to PR-3 group    Shinshi et al. (1987)
                   Unknown function; cloned by ± screen of cDNA
                   library from secondary leaves of TMV-infected plants
        Acidic peroxidase   Extracellular; lignin-forming
        Alexander et al. (manuscript in
        preparation)

        Lagrimini and Rothstein (1987)
1

2

3

4

5
permeability may be the trigger to most, if not all, the wounding responses.  However, the

difficult problem of proving that the cytoplasmic Ca2+ change is the first event in O3 injury

remains.

      Some wound- and pathogen-induced genes that are activated or repressed in Arabidopsis

thalania are found with DNA arrays (Cheong et al., 2002). While these responses may not be

uniform for all plants, they suggest the possibility of wide-ranging gene changes that may occur

with a simple wound and that those changes are wide-ranging and diverse.  As an example, these

responses are related to hormonal responses that are related to jasmonic acid, ET, and auxin
       January 2005
                                              9-66
DRAFT-DO NOT QUOTE OR CITE

-------
 1      pathways; signal transduction responses; and transcription factors for a variety of pathways.
 2      The involvement of ET in wounding and pathogen attacks is discussed in Section 9.4.3.2.
 3
 4      9.3.4.1  Peroxidases
 5           Increases in cytosolic and apoplastic peroxidase activity in response to O3 are often
 6      observed, but the reasons and outcomes of these changes have yet to be fully explained.
 7      Increased activity is frequently correlated with O3 injury. Dass and Weaver (1972) observed that
 8      increases in peroxidase after O3 injury was similar to that observed for plant infection by a virus.
 9      Tingey et al. (1975) observed a 35% decrease in peroxidase activity immediately following O3
10      exposure; however, within 24 to 48 h, activity had increased significantly and was above control
11      level and remained there throughout the remainder of the study.  Dijak and Ormrod (1982) also
12      observed increases in peroxidase activity when two O3-sensitive and two O3-resistant varieties of
13      garden peas (Pisum sativum) were exposed to O3. Peroxidase activity was not related to cultivar
14      sensitivity nor to visible injury. Unfortunately, there are many peroxidases  (Birecka et al., 1976)
15      therefore, any general increase is not specific. In ET-treated leaves, peroxidase reaction
16      products were found between the plasma membrane and the cell wall, suggesting that ET itself
17      could induce peroxidase activity (Abeles et al., 1989a,b; Birecka et al., 1976).
18           At the same time, others examined peroxidase reactions in general  and found two types of
19      peroxidases (designated as acidic or anionic and basic or cationic, EC 1.11.1.7, but also listed as
20      EC 1.14.18.1).  Many types of peroxidases are located in diverse organelles, and each seems to
21      be activated by different conditions (e.g., pH for anionic and cation types and substrates such as
22      guaiacol, syringaldazine, and ascorbate). Peroxidases belong to at least two groups, which
23      catalyze two separate reactions: (1) the reaction of H2O2 with ascorbate  to form DHA, discussed
24      earlier (Thorn and Maretzki, 1985), which is regenerated by plasma membrane electron transport
25      using a dehydrogenase (Gross and Janse, 1977) that is now believed to be a malate/oxaloacetate
26      shuttle through the membrane coupled to a NAD(P)H-cytochrome-b-reductoxidase; and
27      (2) the reaction with coniferyl alcohol (from phenylalanine through phenylalanine ammonia
28      lyase) to form lignin within the wall. The anionic peroxidase thought to be  involved with
29      lignification is within the cell wall (Buchanan et al., 2000; Taiz and Zeiger,  2002).  Some basic
30      peroxidases are maintained within the cell, while some are external to the cell. After wounding
31      (Lagrimini and Rothstein, 1987; Gasper et al., 1985), some basic peroxidases  can be activated by

        January 2005                              9-67        DRAFT-DO NOT QUOTE OR CITE

-------
 1      processes leading to the synthesis of stress ET (Yang and Hoffman, 1984) and/or by excess Ca2+
 2      (Gasper et al., 1985). Elicitor treatment of plants change a series of peroxidases, some of which
 3      are similar to those seen in O3-induced changes (see Table 9-9).
 4           The formation of lignin is due to the phenylpropanoid metabolism (Buchanan et al., 2000).
 5      Tyrosine and phenylalanine are converted to cinnamic and/>-coumaric acid, which are in turn
 6      converted to/>-coumaryl, coniferyl, and sinapyl alcohols, and then into lignins.  Hence, the
 7      peroxidase activity is often measured by one of these substrates (Espelie et al., 1986; Gasper
 8      et al., 1985). However, it is questionable whether apoplastic peroxidase activity is limiting for
 9      lignification; laccases have a prominent role as well.  Also, availability of monolignols is critical
10      for core lignin formation, and  it is unclear whether levels of these metabolites change in
11      response to O3.  Studies by Booker (Booker et al., 1991, 1996; Booker and Miller,  1998; Booker,
12      2000) indicated that O3 did not increase core lignin concentrations in foliage of loblolly pine,
13      soybean, or cotton; although levels of phenolic polymers and cell wall-bound phenolics were
14      elevated in soybean.  Increased phenolic polymers appear to  be lignin in acid-insoluble lignin
15      assays and may well be responsible, along with polyphenol oxidase, for the stippling injury
16      observed in O3-treated plants.  Cell wall function implies the transport of peroxidase molecules
17      out of the cell and,  most likely, the regulation of their activities within the wall space. These
18      extracellular peroxidases may be observed by vacuum infiltration of buffer into leaf air spaces
19      and subsequent centrifugation of the tissues to remove the buffer with the apoplastic enzymes,
20      that wash out (Castillo and Greppin,  1986). However exposure to O3 induces important changes
21      in the plant.  For example, extracellular peroxidase activity in Sedum album leaves increased
22      nearly 3-fold over that in the control plants after 2-h exposure to 0.40 ppm (Castillo et al, 1984).
23      This O3-induced increase of extracellular peroxidase appears to be under the control of Ca2+
24      (Castillo et al., 1984; Heath and Castillo, 1987). Initially, no effect on the anionic activity as
25      measured with syringaldazine (specific electron donor for lignifying peroxidases) was observed,
26      yet 21 hours later, the anionic  peroxidase activity was increased, whereas the cationic (ascorbate
27      measured) peroxidase activity was decreased in O3-treated plants. This suggests an immediate
28      response (ascorbate peroxidase activation) and a secondary response that activates  the lignifying
29      peroxidase via gene activation.
30           The rapid response of cationic peroxidase after O3 exposure may not result from de novo
31      protein synthesis but from the secretion and direct activation by Ca2+ ions of enzyme molecules

        January 2005                              9-68        DRAFT-DO NOT QUOTE OR CITE

-------
 1      already present in the tissue. Cationic peroxidases might attack the peroxides and, in this
 2      manner, act as a detoxifying agent with ascorbate as the substrate in the apoplasm. The effect of
 3      Ca2+ upon peroxidase activity is stronger at low H2O2 concentrations (Penel, 1986). Thus, one
 4      can imagine that, when the H2O2 concentration is low, this peroxidase activation would have a
 5      greater in vivo importance. Furthermore, the secretion of cationic peroxidases into the free
 6      spaces as a result of O3 treatment is accompanied by a simultaneous release of at least one of its
 7      natural substrates (ascorbic acid); this cationic peroxidase exhibits a much higher  affinity
 8      towards ascorbate (up to 6-fold) than the anionic isozyme (Castillo  and Greppin, 1986).
 9
10      9.3.4.2  Jasmonic Acid and Salicylic Acid
11           Salicylic acid (SA) and jasmonic acid (JA) are considered to be regulators of the plant
12      defense response (Figure 9-13) (Buchanan et al., 2000). They tend  to respond more slowly than
13      ET, causing widespread effects in the plant tissues. Both seem to be heavily involved in
14      responses of the plant to O3, once again linking the pathogen/wounding defense to O3-induced
15      injury; however, their roles are far from clear.
16           One of the lipoxidase isoforms is activated by pathogen infection (POTLX-3) within 6 h
17      and accumulates for a week (Kolomiets et al., 2000). This enzyme  is the first stage of the JA
18      pathway which leads to 13-hydroperoxide linolenic acid (HPOT) which is converted either to
19      allene oxide through AOS or to C6 aldehydes through hydroperoxide lyase.  These aldehydes act
20      as signaling agents via systemin (Sivasankar et al., 2000) or volatile odiferous compounds
21      (oxylipins) that have been implicated as antimicrobial toxins (Froehlich et al., 2001).
22      Interestingly, these compounds seem to target the chloroplast envelop where they  interact with
23      its metabolism. As FIPOT and AOs are both implicated in plant defense and are activated by O3,
24      these interactions may be related to how chloroplast enzymes and their mRNAs are involved in
25      O3-induced  injury.
26
27      9.3.4.3  Stress-Induced Alterations in Gene Expression
28           Early  studies addressed the qualitative and quantitative effects of O3 on protein metabolism
29      (Harris and  Bailey-Serres, 1994).  Subsequent reports suggested that the physiologic and
30      metabolic consequences of exposure to O3 was, in part, mediated by increased gene expression.
31      A summary of the gene-linked changes in proteins induced by SAR may be seen in Table 9-9.

        January 2005                              9-69        DRAFT-DO NOT QUOTE OR CITE

-------

   Phospholipid-
     Ethylene—
Salicylic Acid-
                                                 -> /AO Synthas^
                  0='
                              JasmonicAcid
                                    (Howe ef a/., 2000)
                           COOH
                                                Allene Oxide
                                                if
                                            Reductase
                                              HOOC
                                                               13-Hydroperoxy-
                                                                       .  A •  i
                                                                 Imolenic Acid
                                                                   (HPOT)
                                                Traumatic Acid
                                                  (Croft etai, 1993)
                                                                  COOH
                                                                       Systemin
       Figure 9-13.  The pathway leading from phospholipids to jasmonic and traumatic acid.
                    The role of lipoxygenase and the production of a hydroperoxyl moiety from
                    the unsaturated fatty acid is clearly demonstrated.  More importantly,
                    several of the enzymes within this pathway have been shown to be activated
                    by oxidative conditions including O3 exposure. The production of both of
                    these acid species could lead to a general global response of a whole plant to
                    the O3 exposure of a single leaf.
       Source: Howe et al. (2000); Croft et al. (1993); Buchanan et al. (2000).
 1     Of particular note are the productions of PR proteins, chitinase, glucanase, and acidic
 2     peroxidases that appear to be common markers used in many O3 studies. A summary of varied
 3     proteins as measured by changes in the mRNA in A. thalcmia induced by O3 exposure is shown
 4     in Table 9-10.  While studies on Arabidopsis thaliana required high concentrations of O3 to
 5     produce a response, the levels reported in most of the studies did not induce visible injury.  The
 6     types of messages induced included glutathione S-transferase, PAL, ACC synthase, SOD, and
 7     some PRs. Slower increases in messages are seen for other PR and SAR-senescence proteins.
 8     Declines in messages were observed for varied chloroplast enzymes, including those for Rubisco
 9     and chlorophyll binding proteins. A few new proteins were found — a casein kinase and three
10     plasma membrane proteins.  It is interesting to note that few messages for "new" proteins were
11     generated by O3 exposure.
       January 2005
                              9-70
DRAFT-DO NOT QUOTE OR CITE

-------
fa
to
O
o



VO
«^j
o
H
6
o
0
H
O
o
0
o
H
W
Table 9-10. Proteins Altered by Ozone as Measured by Molecular Biological Techniques as mRNA Level or
Other Gene Activity Rather than Enzyme Activity
Exposure
150/300 ppb for 6 h
daily
300 ppb for 6 h daily
200/500/1 000 ppb
for2h

350 ppb for 1-6 h
300 ppb for 6 h
1 50 ppm for 6
h/8 and 14 days
160 ppb for 3-72 h
(a) 250 ppb for 8 h;
(b) 250 ppb for 2 h;
(c) 175 ppb for
8 h/4 days





Identified
proteins fast Slow increase
Physiological events increase response response
Leaf curling; GST, PAL Pxase, SOD
reduced growth
10 bands of 10RNA
Wilting (8 h); premature
senescence

Ethylene production; ACS-6
downward curvature;
water logging
Necrosis in NahG and Chi SOD. Chi GPX
Cvi-0 (accumulating SA), cytAPX, GST1
not in Col-0
Downward rolling of leaf; BCB, ERD 1 ,
early senescence SAG21
Early senescence GST1,VSP2; MT1
(a) little chlorosis or lesions; GSTApx, PAT1
(c) growth retardation CuZn-SOD





Examined, Unknown
Decline but no change proteins Reference
CAT, LOX1 Sharma
(1994)
AtOZIl » casein Sharma
kinasell (1995)
3 plasma Tokarska-
membrane Schlattner
proteins: 75-, 45-, (1997)
35-kDapeptides
ACS-1,-2, -4, -5 Vahala
(1998)
Cab mRNA, Rao (1999)
cyto SOD,
chlGR
Cab, rbcS Atgsr2, MT1 , Miller (1 999)
SAG 12, SAG 13,
SAG 19, SAG 20
CCH Mira (2002)
Fe-SODl GR, Conklin
cab,rbs (1995)






-------
to
o
o
             Table 9-10 (cont'd).  Proteins Altered by Ozone as Measured by Molecular Biological Techniques as mRNA Level or
                                                 Other Gene Activity Rather than Enzyme Activity
Exposure Physiological events
200 ppb for 24 h
Identified
proteins fast
increase response
PR-l,PR-2a,PR-5
AtEDSl,AtGSTl,
AtGST2
Slow increase
response Decline
PR-3b, PR-4
Examined, Unknown
but no change proteins
LOX2,AtOZIl,
PAL, Lhcb,
PAT1,HSP
Reference
Matsuyama
(2002)
         250 ppb for 6 h
                   Lesion initiated on margin
                   and spread inward
                                        rcdl, on
                                        chromosome 1,     Overmyer
                                        single Mendelian    (2000)
                                        trait
to
Abbreviations used in Tables 5 and 6.
GST = Glutathionine synthase
PAL = Phenylalanine ligase
PR-1 =
Pxase = Peroxidase
CAT  = Catalase
LOX1 = Lipoxygenase
ACS-6 = ACC synthatase
SOD = Superoxide dismutase
CuZn-SOD = cyto SOD
Fe-SODl = Chi SOD
cytAPX = Ascorbate peroxidase
Chi GPX = Glutathione peroxidase
Cab mRNA = Chlorophyll a/b binding protein
chl GR  = Gluthatione reductase
BCB  =
ERD1 = Ethylene response
SAG21  = Senescence
rbcS  =  Rubisco small subunit
MT1  =  Mitochondria
H
6
o
o
H
O
o
HH
H
W

-------
 1           The working hypothesis is that O3, which is not eliminated by antioxidants in the cell wall,
 2      alters the properties of the plasma membrane.  Specific polypeptides, indicative of these
 3      antioxidants, are induced.  If specific receptor molecules or channels on the membrane are
 4      affected, the ionic balance within the cytoplasm is changed, leading to altered transcription or
 5      translation of the genes controlling those and other types of polypeptides. Once this membrane
 6      disruption occurs, the cell must mobilize repair systems to overcome the injury. Thus, carbon
 7      and energy sources once destined for productivity, must be used in repair processes.  Some of
 8      these repairs are thought to result from the induction of specific genes.  Photosynthesis is
 9      inhibited by direct inhibition of some of the enzymes, through byproducts of O3 attack or by
10      altered ionic balance. At the very least, the decrease in photosynthesis is a result of an
11      O3-induced decrease in rbcS mRNA.
12
13      9.3.5  Primary Assimilation by Photosynthesis
14      9.3.5.1  Photooxidation:  Light Reactions
15           Photooxidation refers to the oxidation of chlorophyll within the light reaction due to an
16      imbalance between light absorption and the CO2 use to produce carbohydrates. It was
17      discovered in the 1920s and studied under the concept of chlorophyll bleaching and photo
18      autooxidation (Asada, 1999; Rabinowitch, 1945). What generally occurs is that electron transfer
19      from H2O to NADPH declines, and a light reaction overload occurs.  The slowdown of electron
20      transfer may also be  due to inhibition of the dark reactions, through the poor use of small
21      molecular weight carbohydrates or a lowered amount of the fundamental  substrate CO2.
22      To counteract these detrimental reactions, a series of "antioxidant" reactions exist, which
23      eliminate the buildup of oxidative intermediates.
24           A lowered CO2 level, which can be caused by stomatal closure (Heath, 1996), blocks the
25      use of reduced plastoquinone (PQH2) in Photosystem II through NADP reduction in
26      Photosystem I (Hankamer et al., 1997). The buildup of PQH2 reduces the amount of QA,
27      resulting in a buildup of P680+|Pheo~ species (the primary photoact).  The inability to reduce this
28      radical leads to injury to the Dx protein (32 kDa) and its fragmentation into 23-, 16-, and 10-kDa
29      fragments (Hankamer et al., 1997).  Ozone exposure of bean plants leads directly to the loss  of
30      this Dj protein (Pino et al., 1995).  The loss of Dj stimulates the production of new Dl (and its
31      mRNA).  Also, the production of the oxidized form of P680 (P680+) is harmful to the plant because

        January 2005                              9-73        DRAFT-DO NOT QUOTE OR CITE

-------
 1      electron flow from water to P680+ is limited, generating a P680T (the triplet form of P680), which is
 2      highly oxidizing and can lead to dangerous reactions. One form of protection is the use of
 3      p-carotene to convert the triplet form back to its normal state; however, that reaction can lead to
 4      the loss of P-carotene.  Without the protection of P-carotene, oxygen with oxidized products to
 5      produces singlet state of O2. This,  in turn, can react with chlorophyll, leading to ring breakage
 6      that, in essence, leads to chlorosis.  These types of reactions do not seem to occur often, but
 7      chlorosis is one form of visible injury, and loss of P-carotene has been reported. Farage et al.
 8      (1991) and Farage and Long (1999) studied these reactions in wheat and concluded that
 9      alterations to the dark reactions were much more common.
10
11      9.3.6  Alteration of Rubisco by Ozone: Dark Reactions
12           A large body of literature shows that O3 exposure induces a decline in Rubisco (Pell et al.,
13      1997). Treatment of a variety  of plants with O3 at near ambient levels results in a loss of
14      Rubisco and of the mRNA coding for both subunits of Rubisco (rbcS, small and rbcL, large).
15      Because Rubisco plays such an important role in the production of carbohydrates (Figure 9-14),
16      any loss may have severe consequences for the plant's productivity.
17           The study by Noormets et al. (2001) used an exposure system of plants in a FACE
18      exposure facility in which areas of ambient (daytime 360 ppm) and ambient with added CO2
19      (560 ppm), with added O3 (97.8 ppb), and with added CO2 and O3, were used.  Two clones of
20      aspen (O3-tolerant and -sensitive) were used in all four cases.  The study recognized that leaf age
21      was important in the variation  and tried to control for it by grouping the data by leaf plastochron
22      index12. These data, taken over many days using the LiCOR 6400, confirmed earlier
23      observations that O3 has the greatest effect on older leaves after causing a decline in assimilation
24      and in conductance.  They also confirmed that the internal CO2 (calculated for within the leaf) is
25      not affected by O3 exposure. Higher levels of CO2 increased the assimilation and lowered the
26      conductance, maintaining the internal to external CO2 ratio identical to that found with the
27      ambient CO2 level, corresponding to the theory of Farquhar et al. (1980).  The level of Rubisco
28      was not measured as frequently as the assimilation and conductance, but the significant (5%)
29      increase in Rubisco was only observed in older leaves for both clones. More to the point was
               12The plastochron index for the leaf measures the leaf age by degree of expansion rather than simply by
        chronological time.

        January 2005                               9-74       DRAFT-DO NOT QUOTE OR CITE

-------
                                           {precursors}
                           {precursors}r


                         DNA —*>  {rbcJi


                                    {products}r
                                                  {products}
                                                                    {Triose-P}
       Figure 9-14.  The production of Rubisco and its Calvin Cycle pathway reactions. Two
                     peptides are used to build Rubisco:  rbcS, the small subunit produced by
                     DNA within the nucleus; and rbcL, the large subunit produced by DNA
                     within the chloroplast itself. Clearly both polypeptides must be closely
                     regulated to produce the enzyme in  a coherent manner.  Furthermore,
                     at least five isoforms of DNA can produce rbcS, each of which is regulated
                     by a different promotor region.
1      that the stomatal limitation13 was not altered by O3 exposure, with or without excess CO2. It is

2      critical to point out that mesophyll conductance is directly linked to internal CO2 level14. So if

3      Ci/Co is constant and gs declines, then gm must likewise decline. If, as it is argued, Rubisco

4      levels are constant or at least increasing, then a regeneration of RuBP must be the cause of the

5      decline in gm.  Farquhar et al. (1980  were more concerned with high levels of CO2 and had little
              13The limitation was defined as the ratio of stomatal resistance to the total resistance (which included the
       operating point of the assimilation (A) verses internal CO2 concentration (Q) curve and the resistance of the
       boundary layer). The operating point of the curve was defined as the internal CO2 level (which is calculated by the
       conductance and assimilation). The resistance of this operating point was calculated as the cotangent of the slope to
       the operating point. Unfortunately, the slope is not a dimensionless parameter but is rather moles of air per area of
       leaf ~s of time and, thus, it is unclear whether the slope changes with added CO2 and O3.

              14Respiration is generally small at saturating A and often is ignored. By transforming {A = gs (C0 - Q)}
       into {A = gs C0 - gj,, C0 = (gs - g,,,) Co} where gm= gs (Q/C0) or the mesophyll conductance in earlier literature.
       January 2005
9-75
DRAFT-DO NOT QUOTE OR CITE

-------
 1      to say about O3 exposure.  One thing is certain however, the use of clones using different aged
 2      leaves is a preferred method for this type of investigation.
 3           The level of carbohydrate within the cell has an effect upon the amount of mRNA for
 4      Rubisco (rbcS).  Experiments by Krapp et al. (1993) indicated that a decline in carbohydraft
 5      levels is probably due to the increased production of control metabolites, such as fructose
 6      2,6-bisphosphate, which can shut down important sugar production pathways.  This report also
 7      leads to a measure of half-time for the decline in rbcS of about 2 days15 when 50 mM glucose is
 8      added to a cell suspension of Chenopodium.  Also, the carbohydrate level was increased by cold
 9      girdling the petioles of tobacco and potato using intact plants. The levels of carbohydrate nearly
10      doubled in 5 days and the level ofrbcS declined rapidly (reaching 25% after 12 h).  A decline in
11      Rubisco followed, but more slowly (with an estimated half-time of about  108 h after a lag of at
12      least 12 hours).  This, of course, is expected;  the level of the enzyme would decline slowly with
13      a lag after a loss of the message.
14           A better estimation of the half-life ofrbcS can be found in the Jiang et al. (1993) study of
15      the destabilization of the message by an antisense message. The wild type rbcS in tobacco had a
16      half-life of about 5 h compared to that in the mutant with the antisense. It was argued that the
17      antisense message increased the degradation of the normal rbcS.  The estimated half-life ofrbcS
18      under O3 fumigation is about 1 hour (Pell et al., 1994). Although comparisons of these diverse
19      systems can not be easily made, the normal half-life ofrbcS may be  closer to 5 to 10 h; and O3
20      fumigation does not simply stop the transcription of DNA, but rather it alters the rate of
21      degradation, either independently of, or simultaneously with, transcription.
22           Williams et al. (1994) developed a correlation between the levels of ABA after water stress
23      in Arabidosis  thaliana leaves and the loss ofrbcS.  Although their data were not quantitative, the
24      level of ABA  had a half-time rise of about 1 to 2 h and the level ofrbcS had a half-life decline of
25      about 2 to 4 hours.  Their work suggests that  drought stress may alter the CO2 metabolism by
26      changing enzyme relationships much more than by merely closing the stomata. If an ABA rise
27      is lowering rbcS, rbcS may not be a good marker of O3 fumigation except under highly
28      controlled conditions.
29
               15The amount of Rubisco drops from an initial 0.12 to a final amount of 0.04 umole/g-FW s in 6 days.

        January 2005                               9-76         DRAFT-DO NOT QUOTE OR CITE

-------
 1      9.3.7  Carbohydrate Transformations and Allocation
 2           The question of whether translocation of the sugars out of the leaf is inhibited by O3
 3      exposure arises because productivity is often dramatically inhibited by O3 fumigation. Though
 4      nearly 35 years have past since Dugger and Ting (1970) investigated the question of sugar
 5      transport within the leaf, the question has since been little studied. Translocation (Cooley and
 6      Manning, 1987) appears to be inhibited, because root functions are impaired by O3 exposure.
 7      Many observed events suggest that while carbon assimilation within the leaf declines,
 8      translocation of carbon is inhibited even more so, because plant growth points are inhibited and
 9      root/shoot ratios are altered (Dugger and Ting, 1970; Gerant et al., 1996; Tjoelker et al., 1995).
10           Many of the experiments with O3 fumigation indicate that O3 exposure decreases the net
11      growth or dry mass of the plant, but the mechanism is poorly understood.  Generally the
12      decrease in assimilation is much less than the decrease in growth, but not always. Under many
13      conditions, the stomata will close partially, decreasing assimilation by a smaller factor.  Only a
14      long exposure, or high levels of exposure for a short time, generate enough decline in Rubisco to
15      make the growth of the plant problematical. No convincing argument has linked the decrease in
16      growth with a small decline in assimilation, either by a conductance- or Rubisco-limitation.
17      Measures of assimilation with crops are frequently done on upper canopy leaves, which are the
18      last leaves to exhibit O3 injury, while leaves deeper in the canopy exhibit injury and early
19      senescence. Crop root growth must be sensitive to these and other O3 effects, because root
20      biomass is often suppressed early by elevated O3.
21           Volin et al. (1998) found O3 exposures statistically decreased leaf area ratio, specific leaf
22      area, leaf weight ratio, and root weight ratio in Populus tremuloides and two C3 grasses
23      (Agropyron smithii and Koeleria cristatd) but not in Quercus rubra and in the C4 grasses
24      Bouteloua curtipendula and Schizachyrium scoparium.  There was no statistically significant
25      change in any species in leaf conductance (4% level decline in K. cristatd) nor in assimilation
26      (although there was a decline in assimilation at the 6% level for P. tremuloides  and a decline at
27      the  1% level in B. curtipendula). They also reported a correlation between growth decline and
28      decreased stomatal conductance among all  species.
29           Trade-offs are made by plants. Birch grown in highly fertilized conditions exhibited a
30      greater leaf turnover when exposed to  O3, in that leaves not only formed faster but abscised
31      faster, presumably due to early senescence; whereas birch grown under poorer fertilized

        January 2005                              9-77        DRAFT-DO NOT QUOTE OR CITE

-------
 1      conditions retained their leaves longer and had a greater respiration rate within those leaves
 2      (Maurer and Matyssek, 1997).  Again, one must be careful in comparing short-term versus long-
 3      term exposures. Grulke et al. (2001) observed that maximum concentrations of carbohydrates in
 4      1-year old needles that had not abscissed due to early senescence declined when subjected to
 5      year-long exposures along an increasing pollution gradient.  Furthermore, the monosaccharide
 6      concentrations in fine roots (along with starch) were largely decreased, suggesting that needle
 7      sugars were limiting, leading to root sugar limitations. However, determination of the total
 8      productivity and detailed balance of carbohydrate was impossible, because these were older,
 9      larger trees and the data were taken over a full growing season. For a shorter-term exposure of
10      9 days, Smeulders et al. (1995) observed that O3 appeared to increase the retention of labeled
11      photosynthates within the needle, and, at higher exposures (400 versus 200 or 0 mg/m3), the total
12      starch within the needle decreased, suggesting that less carbohydrate was produced within the
13      cell or perhaps that it was in compounds not measured.
14           Studies with Pima cotton, aspen (Populus spp.) and bean seedlings (Phaseolus vulgaris)
15      indicate that acute O3 exposures inhibit export of the current assimilate that provides
16      carbohydrates to the roots from source leaves  of cotton as well as recent assimilate from the
17      older leaves of aspen and bean (Grantz and Yang, 2000).  Grantz and Yuan (2000) attempted to
18      distinguish between potential mechanisms of O3 phytotoxicity operating at the level of the whole
19      plant.  Four hypothesis were tested by fumigating cotton:  (1) O3 exposure reduces leaf pools of
20      soluble sugars; (2) pruning leaf area and reducing source strength to match that of O3-treated
21      plants reproduces O3 effects; (3) pruning lower leaf area more closely reproduces O3 effects than
22      pruning the upper leaf area; and (4) manipulating plant age and, thereby, plant size to match
23      O3-treated plants reproduces O3 effects.  All were shown to be incorrect. Under each  of the
24      above conditions, Grantz and Yang (2000) reduced the amount of foliage to match that caused
25      by O3 injury. While the treatments reduced the biomass and leaf area, they did not alter biomass
26      allocation nor root function. They concluded that a simple loss of foliage does not induce the
27      changes in translocation to the roots to the same extent as does O3 injury.
28           This finding by Grantz and Yang (2000) is important in that it suggests that O3 can trigger
29      a plant-wide response that may be linked to alterations in signal transduction and the generation
30      of whole plant signals.  Stitt (1996) suggested that"... allocation is regulated by long-distance
31      signals that act to influence growth of selected sinks and to modify the delivery of resources to

        January 2005                               9-78        DRAFT-DO NOT QUOTE OR CITE

-------
 1      these sinks in parallel."  Cooley and Manning (1987), citing Me Laughlin and McConathy
 2      (1983), suggested three possible ways that O3 fumigation might alter translocation:
 3      (1) malfunction of the phloem loading process, (2) increased allocation to leaf injury repair, and
 4      (3) an altered balance between the leaf and sinks caused by reduced carbon fixation and a greater
 5      demand for assimilate in the leaf.
 6           Ethylene has been shown to reverse this sugar inhibition of development and to be
 7      antagonistic to the ABA effect (Finkelstein and Gibson, 2002).  However, these effects depend
 8      greatly upon the developmental  stage of the plant.  Thus, the balance of the effectors (sugars,
 9      ABA, and ET) may interact to generate the variation observed in the O3-induced productivity
10      decline. For example, O3 fumigation can induce a shift in the carbon transfer between roots and
11      shoot; and this shift can be amplified by mild drought (Gerant et al., 1996). Furthermore a
12      regulation of source-sink relations with the defense responses induced by elicitors was observed
13      by wounding the leaves of Chenopodium rubrum.  Ethylene appears to be able to repress the
14      expression of extracellular invertase, which is critical for control and down-loading of sucrose
15      derived from the translocational stream (Roitsch, 1999; Lindow et al., 1996).  In addition, the
16      development of Arabidopsis at high concentrations of glucose or sucrose is arrested by
17      increasing the ABA level (Coruzzi  and Zhou, 2001).
18           Clearly more work is needed  on the interactions between assimilation, translocation, and
19      source/sink relations with  O3 exposure.  In these interactions, one must be aware of the
20      developmental age of the plants and their phytohormonal status.
21
22      9.3.7.1  Lipid Synthesis
23           Heath (1984) summarized several early reports of O3-exposure induced  lipid alterations.
24      Most concerned the production of MDA as a measure of lipid oxidation as well as the loss of
25      unsaturated fatty acids. However, a series of experiments by Sakaki and  coworkers concentrated
26      on one type of fumigation system and one metabolic pathway.  This literature provides the best,
27      most complete story with regard to lipid metabolism and O3 fumigation and indicators that O3
28      injures cellular membrane systems  via lipid destruction.
29           Sakaki and coworkers used spinach, which is a sensitive plant but which has not been
30      much evaluated with respect to O3 fumigation. While the O3 level was high (0.5 ppm), enough
31      work has been done to be able to "tease apart" what is happening.  The first paper showed that

        January 2005                             9-79       DRAFT-DO NOT QUOTE OR CITE

-------
 1      chlorophyll bleaching does not begin until the plants have been exposed to O3 for over 10 h,
 2      whereas some MDA production begins with as little as 6 h exposure (Sakaki et al., 1983).
 3      Consistent production of MDA, indicative of gross disruptions, occurred only after 8 h exposure
 4      (Sakaki et al., 1985), within the timescale when chlorophyll and carotenoid levels began to
 5      decline.  Concurrently, the total fatty acid (FA) level decreased from -481 to 358 nmol/cm2 as
 6      the MDA level increased from 0.6 to 2.4 nmol/cm2, indicating FA peroxidation (Sakati et al.,
 7      1985).
 8           Sakaki  et al. (1983) also studied development of changes by cutting disks from exposed
 9      leaves and floating them on water solutions for varied time periods (up to 24 h). This permitted
10      feeding experiments to be done easily, whereas the cutting gives rise to an additional wound
11      response and eliminates metabolite movement to and from other portions of the plant. The
12      floating experiments indicated that, after exposure, scavengers of singlet oxygen (JO2), such as
13      D2O, and of hydroxyl radicals, such as benzoate and formate, have no effect on development of
14      the MDA response after 8 h of in vivo fumigation, while scavengers of (O2 ), such as tiron and
15      ascorbate, lowered the amount of MDA formed.  By measuring metabolites immediately  after
16      cessation of fumigation, they were able to show that ascorbate loss began with the onset of
17      fumigation, as did SOD loss.  A lag time associated with the production of DHA suggested that
18      the reaction of ascorbate with fumigation did not immediately produce the oxidation product.
19      The first 4 h of exposure yielded 30 nmole/cm2 of ascorbate loss with  5 nmole/cm2 of DHA
20      production, whereas the second 4 h of exposure yielded 20 nmole/cm2 of ascorbate loss with
21      20 nmole/cm2 of DHA production.
22           Nouchi and Toyama (1988) exposed Japanese morning glory (Ipomea nil) and kidney bean
23      (Phaseolus vulgaris) to 0.15 ppm O3 for 8 h.  Under these conditions, little visible injury  was
24      found with up to 4 h exposures, while injury increased by -50% after 8 h of exposure.  Morning
25      glory produced more MDA than kidney bean, which produced the same as the zero-time  control.
26      Morning glory also demonstrated a slight (5%) drop in MDGD (monogalactosyldiacylcerol),
27      with increases in PC (phosphatidylcholine), PG (phosohatidylglycerol), PI (phosphatidylinositol),
28      and PE (phosphatidyelthanolamine) after 4 h. Twenty-four hours later, the drop in MGDG
29      (mongalactosyldiacyglyerol) was much larger and was thought to be related to an inhibition of
30      UDP-galactose galactotransferase due to a rise in free fatty acids (FFAs) in the chloroplast. Note
        January 2005                              9-80        DRAFT-DO NOT QUOTE OR CITE

-------
 1      that the two distinct timescales involved in O3 fumigation, immediately post-fumigation and a
 2      day or so later, allows for comparison after plant metabolism responds to the fumigation event.
 3           The pathway for the formation of MGDG and DGDG (digalactosyldiacylglcerol) is located
 4      on the chloroplast envelope. Diacylglycerol (DG) arrives from either the endomembrane system
 5      or the stroma and the enzyme UDP-Galactose: 1,2-diacylglycerol galactosyltransferase (UDGT)
 6      forms MGDG with galactose from UDP-galactose. Sakaki et al. (1990) suggested that the
 7      O3-induced inhibition of UDGT was due to a release of FFAs from within the chloroplast. These
 8      FFAs are inhibitory to UDGT, but not to GGGT, which is stimulated by high concentrations of
 9      Mg2+ (Sakaki et al., 1990).  The Sakaki et al. (1990) data indicate that the measured activities of
10      both enzymes isolated after fumigation has indicated that in vivo are not affected by O3
11      fumigation in vivo. Both enzymes have sensitivity sulfhydryls, and both are located on  the
12      envelope. Ozone, if it reaches those sulfhydryls, should inhibit these enzymes; yet inhibition
13      was not seen.
14           It has been thought for years that tocopherols functioned as antioxidants in biological
15      systems (Tappel,  1972). Hausladen et al. (1990) examined the role of antioxidants in red spruce
16      by following seasonal changes.  They fit the level of tocopherol within the needles (/g-FW) to
17      the time of the year and found little change (fit as level = A + Bt + Ct2).  From this empirical fit,
18      they found that the constant A was lower with higher levels of O3.  The seasonal variation
19      coefficients, B and C, were also lower, suggesting year long lows tocopherol levels.  Variation
20      with the season is not particularly surprising, given that phytochrome  action may be linked to
21      tocopherol biosynthesis (Lichtenthaler, 1977). Hansladen et  al. (1990) reported a significant
22      (p < 0.05) trend in the difference between the high and low level of treatment; although  there
23      was no discussion of why it occurred or what it meant in relation to metabolism. Their major
24      conclusion was that the antioxidant changes due to O3 exposure may decrease cold hardiness.
25           Sterols, believed to act as membrane stabilizers,  have been investigated by several groups
26      with mixed results.  Tomlinson and Rich (1971), who  exposed common bean at 0.25 ppm for
27      3  h, and Grunwald and Endress (1985), who exposed soybean at 0.07  ppm for 6 h for 48 days,
28      reported an increase in free sterols and a decline in esterified sterols. However, Trevanthen et al.
29      (1979) exposed tobacco at 0.3 ppm for 6 h and reported opposite results. None of these
30      investigators believed that O3 had attacked the sterols  directly, instead, they believed that these
31      changes involved metabolism and membrane stability. If O3  induced a metabolic shift that

        January 2005                              9-81        DRAFT-DO NOT QUOTE OR CITE

-------
 1
 2
 3
 4
 5
 6
 7
 8
 9
10
11
12
13
disturbed the polar lipid to sterol balance, membrane reactions to other stressors, such as cold
tolerance, would certainly also be affected, perhaps detrimentally.

9.3.8  Role of Age and Size Influencing Response to Ozone
     Clearly many changes occur with O3 exposure can be observed within hours, or perhaps
days, of the exposure.  This document has argued that many of those events are connected with
wounding and elicitor-induced changes in gene expression, but those are relatively fast acting
changes (a timescale of tens of hours).  Two other effects due to O3 take longer to occur and tend
to become most obvious under long periods of low-O3 concentrations. These have been linked
to senescence or some other physiological response very closely linked to senescence.  These
two responses, separated by a time sequence, are shown diagrammatically in Figure 9-15.
                                              03
                                                  M— Ascorbate •<—
                                           "First Event"
                                       {Membrane Alteration}  U-
                                      i ?
                                            	  ROS	
                                                      I
                                            Hypersensitivity Reactions
                                Accelerated
                                Senescence
       Figure 9-15.  Linkage of senescence with hypersensitivity reactions and first event of O3
                     attack of plants.
 1           The understanding of how O3 affects long-term growth and resistance to other biotic and
 2     abiotic insults in long-lived trees is unclear.  Often, the conditions to which a tree is subjected to
 3     in one year will affect the response of that tree in the next year.  This has been called "memory
 4     effect", although the term "carry-over" is preferred.  In other words, a condition in an earlier
 5     year sets the stage for a reaction in the next year; thereby giving a "cause-effect" scenario.
       January 2005
                                          9-82
DRAFT-DO NOT QUOTE OR CITE

-------
 1           In perennial plant species, growth affected by a reduction in storage carbohydrates may
 2      result in the limitation of growth the following year (carry-over effects).  Carry-over effects have
 3      been documented in the growth of tree seedlings (Hogsett et al., 1989; Sasek et al., 1991; Temple
 4      et al., 1993; U.S. Environmental Protection Agency, 1996) and in roots (Andersen et al., 1991;
 5      U.S.  Environmental Protection Agency, 1996). Accumulation of the carry-over effects over time
 6      will affect survival  and reproduction. Data on the cumulative effects of multiple years of O3
 7      exposures have been,  for the most part, the result of 2- to 3-year seedling studies. The difficulty
 8      of experimentally exposing large trees to O3 has lead to the tacit assumption that seedling
 9      response to O3 is a good predictor of large-tree response to O3 (U.S. Environmental Protection
10      Agency, 1996).
11           The carry-over effects of O3 exposures as observed in tree seedlings cited above by Hogsett
12      et al. (1989) have been termed "memory effects" by Langebartels et al. (1997) and proposed by
13      Schmieden and Wild (1995) to explain the sensitivity  of spruce seedlings to frost in the winter
14      after having been exposed to O3 during the previous summer. Norway spruce exposed to 80 ppb
15      for a whole growing season, demonstrated visible injury symptoms the following year when the
16      new needle flush appeared (Langebartels et al., 1997). Additional studies using Norway spruce
17      and Scots pine seedlings have showed similarly delayed responses following O3 exposures.
18      Carry-over symptoms were noted to develop at different times of the year, depending  on the
19      species of seedling  exposed: in early spring for Norway spruce, and in early autumn for Scots
20      pine  (Lange et al., 1989).  Visible effects of O3 exposures on spruce and pine may develop after a
21      substantial delay during the "sensitive" periods of the year when chlorophyll and needle loss
22      normally occur.  Norway spruce and Scots pine differ in their sensitive periods because of the
23      different needle  classes normally remaining on the tree (Langebartels et al., (1997).
24           Nutrient status of the tree during the over-wintering phase of its life (Schmieden and Wild,
25      1995) and chronic exposure to ambient O3 (less severe with fewer peaks of very high levels)
26      induce (1) mineral nutrient deficiency; (2) alterations  of normal metabolism, including
27      translocation and allocation of carbohydrates and probably nitrogen; and (3) disturbance of
28      normal transpiration and diurnal cycling, leading to water stress.  This condition, termed
29      "Montane yellowing", appears to be related to nutrient deficiencies rather than senescence
30      (although early loss of leaves and needles does occur). While generalized low nutrient
31      concentrations may not occur within the foliage, localized deficiencies may. However, they are

        January 2005                               9-83        DRAFT-DO NOT QUOTE OR CITE

-------
 1     hard to observe or prove without a great deal of work involving all portions of a tree and without

 2     a general hypothesis of what is occurring.
 3

 4     9.3.9  Summary

 5           As the understanding of wounding responses of plants and more genome details and varied

 6     plant mutants become available, the cellular and physiological responses of plants to O3

 7     exposures are slowly becoming clearer. However, more studies are needed on a larger variety of

 8     species.  Nevertheless, several key findings and conclusions can be highlighted:

 9           (1)  The entrance of O3 into the leaf through the stomata remains the critical step in O3
                  sensitivity. Not only does O3 modify the opening  of the stomata, usually closing it
                  partially, but O3 also appears to alter the response  of stomata to other stressful
                  situations, including a lowering of water potential and ABA responses.  The
                  concentration of O3 within the leaf is not the same as the external concentration due
                  to reactions within the leaf, but it is not "zero".

10           (2)  The initial reactions of O3 within  the leaf are still unclear, but the involvement of
                  H2O2 is clearly indicated. The detection of possible products by EPR spectroscopy
                  has progressed, but has not reached the point where any products can be identified.
                  Nonetheless, reaction of O3 (or its product) with ascorbate and possibly  other
                  antioxidants present in the apoplastic space of the mesophyll cells  is clear and
                  serves to lower the amount of O3  or product available to alter the plasma membrane
                  of the cells.

11           (3)  The initial sites of membrane reactions seem to involve transport properties and,
                  possibly, the external signal transducer molecules. The alteration and mechanism of
                  the alteration of the varied carriers of K+and Ca2+  is far from clear, but it would
                  seem that one of the primary triggers of O3-induced cell responses  is a change in
                  internal Ca2+ levels.

12           (4)  The primary set of metabolic reactions that O3 triggers now clearly includes those
                  typical of "wounding" responses  generated by cutting of the leaf or by
                  pathogen/insect attack. Again, this seems to be due to a rise in cytoplasmic Ca2+
                  levels.  Ethylene release and alteration of peroxidases and PAL activities,  as well as
                  activation of many wound-derived genes, seem to be linked to some of the primary
                  reactions.

13           (5)  The alteration of normal metabolism due to wounding has effects outside of the
                  cytoplasm. What  effects are due  to the "spreading of the problem" to other cellular
                  organelles is less clear. One of the secondary reactions is linked to an activation of
                  a senescence response. The loss of Rubisco and its messenger RNA is linked to an
                  early senescence or a speeding up of normal development leading to senescence.
       January 2005                               9-84        DRAFT-DO NOT QUOTE OR CITE

-------
                  The loss of photosynthetic capacity is linked to the lowered productivity of plants,
                  and problems with efficient translocation are indicated, although photosynthesis and
                  translocation still occur at a reasonable rate.  The loss of productivity is not yet
                  clearly explained.

14           It is important to note that the dramatic strides in understanding the genetic makeup of
15      plants, gene control, and signal transduction/control over the last few years will likely accelerate
16      in the future. That understanding will translate into better models of the hypotheses listed above
17      and more detailed schemes of how O3 alters much of basic plant metabolism. Thus, while
18      understanding of how O3 interacts with the plant at a cellular level has dramatically improved,
19      the translation of those mechanisms into how O3 is involved with altered cell metabolism, with
20      whole plant productivity, and with other physiological facts remained to be more fully
21      elucidated.
22
23
24      9.4  MODIFICATION OF FUNCTIONAL AND GROWTH RESPONSES
25      9.4.1  Introduction
26           The responses of plants to any air pollutant may be significantly influenced by a wide
27      range of biological, chemical and physical factors. A plant's genetic make-up is an important
28      inherent biological determinant of its response, but response can also be modified by other
29      biological agents such as disease-causing organisms, insects and other pests, and by other higher
30      plant species with which it may be competing for resources.  Chemical factors that may
31      influence response range from mineral nutrients  obtained from the soil to other air pollutants and
32      agricultural chemicals. Physical factors that may influence response include light, temperature
33      and the availability of moisture, which are components of climate and climate change.
34           Some environmental factors are capable of being controlled, to some degree, by man,
35      while others are not.  The biological factors (pests, diseases, symbioses and competition) are
36      partly controllable in agriculture but much less so (if at all) in natural ecosystems. It is possible
37      to control agricultural soil fertility and the use of agricultural chemicals, as well as to exercise
38      some control over the supply of water and airborne chemical factors.  In contrast, the physical
39      factors, i.e., light and temperature, are uncontrolled in the field even though they may be
40      controllable in specialized situations such as greenhouses or shade houses.

        January 2005                              9-85        DRAFT-DO NOT QUOTE OR CITE

-------
 1           The impacts of these various factors on plant response to O3 and other oxidants were
 2      extensively reviewed in the 1996 O3 AQCD (U.S. Environmental Protection Agency, 1996).
 3      It was noted in that document that, since any combination of these factors may come into play at
 4      some time in a plant's life history,  "response will be dictated by the plant's present and past
 5      environmental milieu, which also includes the temporal pattern of exposure and the plant's stage
 6      of development."  That document also stressed that both the impact of environmental factors on
 7      response to oxidants and the corollary effects of oxidants on responses to environmental  factors
 8      have to be considered in determining the impact of oxidants on vegetation in the field. The
 9      variability observed in plant responses to defined exposures to O3, particularly under field
10      conditions, is a consequence of the influences of genetics and the range of  environmental
11      variables.
12           In view of the large number of factors to be considered and given that the purpose of this
13      document is to support the review  of the O3 NAAQS, including standards to protect vegetation,
14      this section focuses mainly on situations in which there is clear evidence that environmental
15      factors truly interact with oxidant effects, i.e., they magnify or diminish the impact of O3 and are
16      not merely additive to it. Conversely, it will cover situations where O3 acts synergistically or
17      antagonistically, but not additively, with effects induced by other factors. It will also emphasize
18      those interactions as a result of which overall plant growth and development, and yield are
19      adversely affected, rather than the  details of interactions at the mechanistic level, unless the latter
20      are deemed to be essential to an understanding of larger scale effects.
21           To facilitate cross-reference,  the present document uses essentially the same subsections as
22      in the  1996 O3 AQCD (U.S. Environmental Protection Agency, 1996). Although light and
23      temperature are components of climate, they are initially reviewed as individual physical factors,
24      even though temperature effects are revisited to some extent in the discussion of interactions
25      with climate change.
26           Few  studies reported since the 1996 document have  systematically investigated
27      quantitative responses to O3 concurrently with other variables. Although the 1996 document
28      cited almost 300 references pertaining to environmental interactions, and the present review cites
29      more than  350 new references, the bulk of the recently published work has continued to be
30      determined by the specific and frequently narrowly focused interests of individual researchers or
31      groups.  Hence, the new findings are scattered and far from uniformly distributed among the

        January 2005                              9-86        DRAFT-DO NOT QUOTE OR CITE

-------
 1      various sub-topics. In some instances there has been little or no research published that adds to
 2      our understanding since the 1996 document. In such cases, the present review is therefore
 3      restricted to summarizing the understanding that was current in 1996.
 4           A few reviews have appeared since the early 1990s dealing with various environmental
 5      interactions, and these are cited in relevant sections below. More general recent reviews are
 6      those of Wellburn (1994); multi-authored volumes edited by Alscher and Wellburn (1994),
 7      Yunus and Iqbal (1996), De Kok and Stulen (1998), and Bell and Treshow (2002); and reports
 8      by the United Nations Environment Programme (UNEP) (1999) and the Intergovernmental
 9      Panel on Climate Change (IPCC) (2001). Several biotic and abiotic interactions involving forest
10      trees are discussed in the review by Johnson et al. (1996b).
11           Although many reports have provided quantitative information on interactive effects, in
12      most cases the information is only descriptive of a specific situation involving only two or three
13      levels of a variable.  While this may be adequate to provide statistical information about the
14      existence of interactions with environmental factors, it does not permit the development of
15      response surfaces or models to show the form that any influence of such factors might take on
16      O3 exposure-response relationships or how O3 might quantitatively influence responses to the
17      factors in question. This, together with the fragmented information available on the effects of
18      most factors, has contributed to the relative  lack of development of simulation models of
19      oxidant-environmental factor interactions. Yet, as noted by Taylor et al. (1994), the large
20      number of variables constrains the assessment of pollution effects by experimentation alone.
21      The only alternative is to use mathematical models to attempt to predict the outcome of different
22      O3 and environmental factor scenarios, building up their complexity in stages. The few models
23      thus far used to investigate O3 stress have been adapted from existing process models of crop or
24      tree growth which include limited numbers of physical or chemical variables (such as
25      temperature,  soil water stress, or nutrient deficiency). Taylor et al. (1994) provide a listing of
26      several simulation models developed for trees at the individual, stand,  and regional levels, and
27      these and many  other models have been critically reviewed by Kickert and Krupa (1991) and
28      Kickert et al. (1999). However, regardless of whether such models are descriptive/empirical or
29      process/mechanistic, their outputs will always be associated with varying degrees of uncertainty
30      and require validation against observable responses wherever possible. Kickert et al.  (1999) also
31      point out that very few of the models that have been described provide risk assessments that

        January 2005                               9-87        DRAFT-DO NOT QUOTE OR CITE

-------
 1      address likelihood, in contrast to consequence assessments that address the magnitudes of
 2      effects. Thus, even though capable simulation models of plant response to O3 involving
 3      complex mixes of many biological, physical and chemical factors may be out of reach at the
 4      present time, the use of newer mathematical approaches such as artificial neural networks
 5      (ANNs) has enabled insightful analyses to be performed in several field studies involving
 6      numerous micrometeorological and other environmental variables (e.g., Balls et al., 1996;
 7      Mills et al., 2000).
 8           Because the ensuing subsections deal with studies of O3 interactions involving an
 9      extremely wide array of biological, physical, and chemical factors in the plant's environment,
10      it is inevitable that many different exposure facilities and regimes have been used in these
11      studies. To provide specific information regarding the O3 exposure concentrations, profiles,
12      hours and days of exposure (as well as the types of systems and facilities used for the exposures)
13      would add a wealth of detail that would do little to assist our understanding of the roles of
14      environment factors in modifying the impact of O3 on vegetation or to facilitate our ability to
15      estimate the magnitudes of any such modifications.  Thus, only experiments in which the
16      exposure levels and regimes were within the bounds of ambient experience in North America
17      are discussed in the ensuing subsections, regardless of the type of exposure profile used. The
18      cutoffs used have been: -200 ppb for peak hourly concentrations or for short-term exposures;
19      -100 ppb for daytime means involving prolonged exposures for several hours; or a doubling of
20      ambient levels, in cases in which enriched exposure levels were a function of ambient levels.
21      Actual details of the exposure regimes and conditions can, of course, be obtained from the
22      original references but are only stated here when any distinction needs to be made between the
23      effects of different exposure levels. Hence, it  should be understood that ensuing statements such
24      as "... it was found that O3 caused ..." should always be read as "... it was found that exposures to
25      O3 (within the range of those that have been measured in ambient air) caused ..."
26
27      9.4.2  Genetics
28           The response of individual plants to O3is affected by several factors including the
29      environment in which it is growing, competition with neighbor plants, ontogeny, and genetics.
30      This section examines the role of genetics in plant response to O3. In addition, major knowledge
31      gaps in the understanding of genetic aspects of O3 response are pointed out.

        January 2005                              9-88        DRAFT-DO NOT QUOTE OR CITE

-------
 1           It is well known that species vary greatly in their responsiveness to O3 (Environmental
 2      Protection Agency, 1996).  This again has been recently demonstrated for grassland species
 3      (Warwick and Taylor,  1995; Pleijel and Danielson, 1997; Grime et al., 1997; Bungener et al.,
 4      1999a, b; Franzaring et al., 2000), wild herbaceous plants (Bergmann et  al., 1999; Danielsson
 5      et al., 1999; Nussbaum et al., 2001), agricultural crops (Renaud et al., 1997; Ollerenshaw et al.,
 6      1999; Benton et al., 2000; Heagle and Stefanski, 2000; Fumagalli et al., 2001; Elagoz et al.,
 7      2002; Nali et al., 2002; Kollner and Krause, 2003), horticultural shrubs and trees (Hormaza
 8      et al., 1996; Findley et al., 1997), and forest trees (Paakkonen et al., 1997; Volin et al., 1998; Pell
 9      et al., 1999; Bortier et al., 2000; Postiglione et al., 2000; Landolt et al., 2000; Zhang et al., 2001;
10      Oksanen and Rousi, 2001; Matsumura, 2001;  Guidi et al., 2001; Momen et al., 2002; Nali et al.,
11      2002; Saitanis and Karandinos, 2002).  These studies have shown a wide range of responses to
12      O3, from growth stimulation by a few species  such as Festuca ovina L. (Pleijel and Danielson,
13      1997) and Silene dioica and Chrysanthemum leucanthemum (Bungener et al., 1999) to
14      significant growth reduction, depending on environmental conditions and exposure dose.
15           While determining the explanation for differences in species sensitivity to O3 remains one
16      of the challenges facing plant biologists (Pell  et al., 1999), a number of hypotheses have been
17      suggested. Reich (1987) proposed that variation in O3 sensitivity could be explained by variation
18      in total uptake of the gas. Others have suggested that (1) fast-growing species are more sensitive
19      than  slower growing ones (Bortier et al., 2000), (2) overall O3 sensitivity may be closely linked
20      to root responses to O3 (Warwick and Taylor,  1995) or (3) the relative ability of species to
21      detoxify O3-generated reactive oxygen free radicals may determine O3 sensitivity (Alscher et al.,
22      1997; Pell et al., 1999). Volin et al.,  (1998) suggest that the relative rate of stomatal
23      conductance and the photosynthesis rate at a given conductance both contribute strongly to
24      determining species sensitivity to O3.  Likely, there is more than one mechanism determining
25      sensitivity, even in a single species.
26           Within a given species, individual genotypes or populations can also vary  significantly in
27      O3 sensitivity (Environmental Protection Agency, 1996). For example, the intraspecific
28      variation in O3 sensitivity for was a factor of two for Phleum prateme (Danielsson et al., 1999)
29      and Trifolium repens L. (Postiglione  et al., 2000). A similar range of intraspecific variations in
30      O3 responses was demonstrated for clonal differences in Betulapendula  by Paakkonen et al.
31      (1997) and Primus serotina (Lee et al., 2002).  These examples of wide ranges within species

        January 2005                              9-89        DRAFT-DO NOT QUOTE OR CITE

-------
 1      responses suffice to show that caution should be taken when ranking species categorically as
 2      having an absolute degree of O3 sensitivity (Davison and Barnes, 1998).
 3
 4      9.4.2.1  Genetic Basis of O3 Sensitivity
 5           Plant response to ozone is determined by genes that are directly related to oxidant stress
 6      and to an unknown number of genes that are not specifically related to oxidants.  The latter
 7      includes genes that control leaf and cell wall thickness, stomatal conductance and the internal
 8      architecture of the air spaces.  Although there is currently a great emphasis on individual
 9      antioxidants that can be manipulated by molecular methods, the challenge is to determine the
10      relative contributions of all of the components to plant response and to understand the interplay
11      between them. Recent studies using molecular biological tools are beginning to increase the
12      understanding of O3 toxicity and differences in O3 sensitivity.
13           While much of the research in developing the understanding of O3 responses has been
14      correlative in nature, recent studies with transgenic plants have begun to positively verify the
15      role of various genes and gene products in O3 tolerance.  Overexpressing MnSOD in chloroplasts
16      increased O3 tolerance in transgenic tobacco plants (Van Camp et al., 1994) provided the first
17      definitive proof of antioxidants key role in O3 tolerance.  Subsequently, Broadbent et al. (1995)
18      showed that overexpression of pea glutathione reductase simultaneously in both chloroplasts and
19      mitochondria of transgenic tobacco enhanced O3 tolerance.  Similarly, increased O3 tolerance to
20      O3-induced foliar necrosis was shown for transgenic tobacco plants Overexpressing the cytosolic
21      Cu/Zn-SOD gene (Pitcher and Zilinskas, 1996). Transgenic tobacco plants expressing antisense
22      RNA for cytosolic ascorbate peroxidase, which reduces ascorbate peroxidase production,
23      showed increased susceptibility to O3 injury suggesting a key role in O3 tolerance for the
24      antioxidant ascorbate peroxidase (Orvar and Ellis, 1997).
25           The consensus among molecular studies of O3 sensitivity is pointing to O3 triggering
26      salicylic acid, ethylene, and jasmonic acid and that the signaling of these molecules determines,
27      in some cases, the O3  susceptibility of plants (Overmyer et al., 2000; Rao and Davis,  1999; Rao
28      et al., 2000; Langerbartels et al., 2002; Moeder et al., 2002; Tamaoki et al., 2003; Vahala et al.,
29      2003a, b).  Increased levels of jasmonic acid production in O3-tolerant compared to O3-sensitive
30      plants has been shown for Arabidopsis (Overmyer et al.,  2000) and Populus (Koch et al., 1998,
31      2000). Blockage of ethylene production by using antisense methods with ACC synthase and

        January 2005                              9-90        DRAFT-DO NOT QUOTE OR CITE

-------
 1      ACC oxidase suggest strongly that ethylene synthesis and perception are required for H2O2
 2      production and cell death following O3 exposure of Lycopersicon esculentum (Moeder et al.,
 3      2002).  Ethylene signaling may have multiple roles in O3 tolerance determination as was
 4      demonstrated recently by Vahala et al., (2003a, b) who found that in Populus tremula x
 5      P. tremuloides hybrid clones differing in O3 sensitivity, ethylene accelerated leaf senescence in
 6      sensitive plants under low O3, but under acute O3,  ethylene seemed to be required for protection
 7      from cell death.
 8           While changing expressions of single antioxidant genes has proven very useful in
 9      identifying possible mechanisms of O3 sensitivity and tolerance (Kuzniak, 2002), it should be
10      noted that some studies of transgenic plants with enhanced antioxidant production have not
11      resulted in increased O3 tolerance (Saji et al., 1997; Torsethaugen et al., 1997; Strohm et al.,
12      1999, 2002). Clearly, ethylene production plays a role in O3 sensitivity but the role of various
13      antioxidants in O3 tolerance regulation are yet to be fully elucidated (Wellburn and Wellburn,
14      1996).  It is unlikely that single genes are responsible for O3 tolerance responses, except in rare
15      exceptions (Engle and Gabelman, 1966).  Regulation of stomatal opening and leaf structure
16      (Bennett et al., 1992; Elegoz and Manning, 2002) are also likely to play key roles in O3 tolerance
17      in plants. Newly developing opportunities to examine simultaneous regulation of larger numbers
18      of genes will likely yield more clarification of genes controlling O3 tolerance (Desikan et al.,
19      2001; Matsuyama et al., 2002).
20           Attempts to demonstrate conclusively changes in antioxidant and protective pigments for
21      O3 sensitive and tolerant  mature trees growing in the field have largely been unsuccessful  (Tausz
22      et al., 1999a, b). However, evidence for antioxidant expression differences contributing to
23      differences in O3 sensitivity of four-year-old Populus tremuloides trees has been found
24      (Wustman et al., 2001).
25
26      9.4.3  Environmental Biological Factors
27           The biological factors within the plant's environment that may influence its response to O3
28      directly or may be influenced to the advantage or disadvantage of plants encompass insects and
29      other animal pests, diseases, weeds and other competing plant species. Although such
30      interactions are ecological in nature, those involving individual pests, plant pathogens, or weeds,
        January 2005                              9-91        DRAFT-DO NOT QUOTE OR CITE

-------
 1      or agricultural crop or forest tree species are considered in this section.  More complex
 2      ecological inter-species interactions are dealt with in Section 9.5.
 3           The different types of biological factors are dealt with separately,  as in the 1996 O3 AQCD
 4      (U.S. Environmental Protection Agency, 1996).  Still, it is important to  recognize certain general
 5      features of relationships of plants with biological components of their environment:
 6        •  Successful infestation or infection involves complex interactions among the target or host
            species, the causal organism and environmental factors.
 7        •  Infestations and infections may co-occur.
 8        •  The successful development and spread of a pest, pathogen or weed require favorable
            environmental factors.
 9        •  Significant losses to crops and forest trees result from pests and pathogens.
10        •  Significant losses to crops and seedling trees result from weed competition.

11           Ozone and other photochemical oxidants may influence the severity of a disease or
12      infestation by a pest or weed, either by direct effects on the causal species, or indirectly by
13      affecting the host, or both. In addition, the interaction between O3, a plant and  a pest, pathogen
14      or weed may influence the response of the target host species to O3.  A perceptive overview of
15      the possible interactions of O3-exposure with insect pests and fungal diseases has been provided
16      by Jones et al. (1994), based on a model system involving two insects and two pathogens
17      affecting cottonwood (Populus deltoides).  Their study also included effects on the
18      decomposition of leaf litter.
19           In contrast to detrimental biological interactions, there are mutually beneficial relationships
20      or symbioses  involving higher plants and bacteria or fungi. These include  (1) the nitrogen-fixing
21      species Rhizobium and Frankia that nodulate the roots of legumes and alder, and (2) the
22      mycorrhizae that infect the roots of many crop and tree species, all of which may be affected by
23      exposure of the host plants to O3.
24           In addition to the interactions involving animal pests, O3 may also have indirect effects on
25      higher herbivorous animals, e.g., livestock, due to O3-induced changes in feed quality.
26
27      9.4.3.1  Oxidant-Plant-Insect Interactions
28           The 1996 O3 AQCD (U.S. Environmental Protection Agency, 1996) stressed the variability
29      in the reported effects of O3 on host plant-insect interactions.  Since relatively few plant-insect

        January 2005                               9-92        DRAFT-DO NOT QUOTE OR CITE

-------
 1      systems have been studied, few consistent patterns of response have emerged, as noted in other
 2      reviews such as those of Colls and Unsworth (1992), Heliovaara and Vaisanen (1993), Whittaker
 3      (1994), Docherty et al. (1997) and, most recently, Fluckiger et al. (2002).
 4           None of the studies reported in the past decade have clarified the situation in terms of
 5      clearly consistent effects.  A 1997 review by Docherty et al. (1997), for example, examined
 6      17 reports of studies of aphid species on a range of hosts and classed the O3 effects on aphid
 7      performance as: 35% positive; 41% negative; and 24% showing no significant effect.
 8      A tabulation of 19 studies by Fluckiger et al. (2002) gave the corresponding figures:  42%,
 9      21%, and 3 7%.
10           Other recent studies with the aphids Schizolachnuspineti and Cinarapinea on  Scots pine
11      (Pinus sylvestris) and Cinarapilicornis on Norway spruce (Picea abies) have also yielded
12      variable results, with them suggesting that O3 enhances aphid density on pine and aphid
13      performance on spruce (Holopainen et al., 1997; Kainulainen et al., 2000a).  In an earlier study
14      with Schizolachnus pineti on Scots pine, Kainulainen et al. (1994) had observed no significant
15      effects of O3-treatment on aphid performance.  However, more recent observations of long-term
16      effects on aphid populations on aspen (Populus tremuloides) exposed to O3 in a FACE system
17      revealed that O3 significantly increased aphid populations and decreased the populations of
18      predatory insects  (Percy et al., 2002).
19           The observations of Brown et al. (1993) and Jackson (1995) led Whittaker (1994) and
20      Brown (1995) to suggest that aphid response was dependent on ambient temperature as well as
21      the dynamics of O3 exposure; growth tended to be stimulated with maximum temperatures below
22      -20 °C but was reduced at higher temperatures. The present situation with plant-aphid responses
23      therefore remains confused and, although numerous suggestions have been offered to explain
24      specific findings, they are difficult to assemble into a coherent picture.
25           Variability has also been  found with the interactions involving chewing insects. For
26      example, Lindroth et al. (1993) reported a small negative O3 effect (8% reduction) on the growth
27      of gypsy moth larvae (Limantra dispaf) on hybrid poplar (Populus tristis x balsamiferd) but no
28      effect when on  sugar maple (Acer saccharum).  Ozone resulted in reduced growth rate of the
29      larvae of the bug Lygus rugulipennis on Scots pine, but enhanced the growth of larvae of the
30      sawfly Gilpiniapallida (Manninen et al., 2000).  Costa et al.  (2001) observed no significant  O3
        January 2005                              9-93        DRAFT-DO NOT QUOTE OR CITE

-------
 1      effects on the growth and fecundity of the Colorado potato beetle (Leptinotarsa decemlincatd)
 2      on potato in greenhouse and field experiments.
 3           Fortin et al. (1997), in a two-year study of forest tent caterpillar (Malacosoma disstrid) on
 4      sugar maple, observed that O3 exposure only increased the growth rate of female larvae in one
 5      year; 4th- and 5th-instar larvae also showed a feeding preference for treated foliage in that year.
 6      However, studies based on open-air exposures of aspen (Populus tremuloides) indicated
 7      O3-enhanced growth of M disstria in terms of pupal weight (Percy et al., 2002) and larval
 8      performance (Kopper and Lindroth, 2003).  Jackson et  al. (2000) observed inconsistency in
 9      studies on the larva of the tobacco hornworm (Manduca sexto) on tobacco (Nicotiana tabacum).
10      In one year, feeding on O3-treated foliage resulted in significantly greater larval weight, whereas
11      in a second year the increase was not statistically significant although survival was increased.
12      Also, oviposition by hornworm moths was increased if ambient O3 levels were increased by 70%
13      but fell back to normal in ambient O3 levels (Jackson et al., 1999).
14           Studies of the two-spotted spider mite (Tetranychus urticae) on white clover (Trifolium
15      repens) and peanut (Arachis hypogeae) by  Heagle et al. (1994) and Hummel et al. (1998)
16      showed that, on peanut and an O3-sensitive clover clone, O3-exposure stimulated mite
17      populations.  The lack of significant effects on mites on the O3-resistant clover clone suggests
18      that the responses were host-mediated.
19           With chewing insects and mites, there therefore appears to be a clearer indication of the
20      likelihood that increased insect performance will result from O3-induced changes in the host
21      plant, but negative effects continue to be reported, indicating that the response is probably also
22      being determined in part by other environmental, genetic, or temporal variables.
23           Reported O3-induced enhancement of attack by bark beetles (Dendroctonus brevicomis) on
24      Ponderosa pine (Pinusponderosd) has been suggested  by Dahlsten et al. (1997) to be due to
25      greater brood development on injured trees, possibly related to decreased numbers of predators
26      and parasitoids. This view gains some support from the observation that O3 adversely affected
27      the searching behavior of a parasitoid, Asobara tabida, for larvae ofDrosophila subobscura
28      which led to fewer parasitized hosts (Gate et al., 1995). Such observations reveal another level
29      of complexity in the O3-plant-insect interrelationship:  O3 may reduce the effectiveness of the
30      natural control of insect pests. The phenomenon is probably related to effects on olfactory cues,
        January 2005                              9-94        DRAFT-DO NOT QUOTE OR CITE

-------
 1      since it was shown by Arndt (1995) that O3 can affect fly behavior by modifying the pheromones
 2      that cause fly aggregation.
 3           These reports focus on the direct or indirect effects on the insect or mite feeding on foliage
 4      previously or currently exposed to O3. They provide little if any information on effects on the
 5      host plant other than qualitative references to the injury caused by the O3-exposure. Enhanced
 6      pest development will ultimately lead to increased adverse effects on the hosts in the long term,
 7      but the only report of an O3-plant-insect interaction directly affecting the host plant in the short
 8      term still appears to be that of Rosen  and Runeckles (1976).  They found that infestation by the
 9      greenhouse whitefly (Trialeurodes vaporariorum) sensitized bean plants (Phaseolus vulgaris) to
10      injury by otherwise non-injurious low levels of O3, leading to premature senescence of the
11      leaves.
12           The overall picture regarding possible O3 effects on plant-insect relations, therefore,
13      continues to be far from clear. Only a few of the very large number of such interactions that
14      may affect crops, forest trees and other natural vegetation have been studied. The trend
15      suggested in the 1996 criteria document that O3 may enhance insect attack has received some
16      support from a few recent studies.  However, the variability noted in most of the studies makes it
17      clear that we are still far from being able to predict the nature of any particular O3-plant-insect
18      interaction or its magnitude or severity.
19
20      9.4.3.2  Oxidant-Plant-Pathogen Interactions
21           Plant diseases are caused by pathogenic organisms, e.g., fungi, bacteria, mycoplasmas,
22      viruses, and nematodes. Ozone impacts on disease are briefly discussed in earlier reviews by
23      Ayres (1991) and Colls and Unsworth (1992) and, more recently, by Fluckiger et al. (2002).
24      Biotic interactions with forest trees have been reviewed by Chappelka and Samuelson (1998);
25      and Sandermann (1996) and Schraudner et al. (1996) have summarized molecular similarities
26      and interrelationships between necrotic O3 injury to leaves and pathogen attack. A few recent
27      publications have added to  our fragmented knowledge of O3-plant-disease interactions and the
28      mechanisms involved, but there appear to have been no reports to date of studies involving
29      mycoplasmal diseases.
30           The 1996 criteria document (U.S.  Environmental Protection Agency, 1996) noted the
31      concept put forward by Dowding, (1988) "that pathogens and pests which can benefit from

        January 2005                             9-95       DRAFT-DO NOT QUOTE OR CITE

-------
 1      damaged host cells and from disordered transport mechanisms are enhanced by pollution insult
 2      to their hosts, whereas those pathogens and other symbionts which require a healthy mature host
 3      for successful invasion are depressed by pollutant stress to their host." The pathogens of the first
 4      type are mostly facultative, necrotrophic, fungal parasites, whereas the second type are largely
 5      obligate biotrophic fungi, bacteria and viruses.  Based on this distinction, the majority of the
 6      cases cited in the 1996 document supported Dowding's view, as have several more recent studies
 7      summarized in Table 9-11.  However, there are also some contradictions.
 8           Most investigations have focused on the incidence and development of disease on plants
 9      previously or concurrently exposed to O3, rather than on the corollary effect of disease on the
10      response to O3. In all of the studies of facultative pathogens and the nematode studies, exposure
11      to O3 tended to result in increased disease severity through increased spore germination or
12      increased fungal growth and development,  although in the case of grey mold (Botrytis cinered)
13      on kidney bean (Phaseolus vulgaris) this was only observed on  an O3-sensitive cultivar
14      inoculated with conidia (Tonneijck, 1994). Using mycelial inoculation, O3 reduced disease
15      development, but no satisfactory explanation was offered to account for the difference in
16      response. With leaf spot, Marssonina tremulae, on hybrid poplar (Populus trichocarpa x
17      balsamifera), low level exposures to O3 also increased disease (in  agreement with theory) but
18      higher levels (200 ppb, 8h per day for 15 days) reduced conidial germination (Beare et al.,
19      1999b).
20           The situation with obligate biotrophic pathogens is less consistent.  The effects on powdery
21      mildew (Sphaerotheca fulginea) on both bottle gourd (Lagenaria siceraria)  and cucumber
22      (Cucumis sativa) resembled the situation with the necrotrophic poplar leaf spot disease
23      (Marssonina tremulae), since low O3 exposures increased disease  severity (in disagreement with
24      theory) although higher levels decreased it. The decreased infection in the pea-powdery mildew
25      (Erysiphe polygon!) situation agrees with theory, but the situations with leaf rust (Melampsora
26      sp.) on poplar (Populus trichocarpa x balsamifera) or aspen (Populus tremuloides) do not.
27      However, these reports are in contrast to earlier reports included in the 1996 criteria document
28      (U.S. Environmental Protection Agency, 1996) of observations with other species of Erysiphe
29      (Tiedemann et al., 1991) and Melampsora (Coleman et al., 1987).  In contrast to the recent report
30      of a synergism with Melampsora on poplar, infections caused by the other biotrophs
31      (Sphaerotheca, Erysiphe, Uromyces)  reduced the severity of injury caused by O3 (in agreement

        January  2005                             9-96        DRAFT-DO NOT QUOTE OR CITE

-------
3
Table 9-11. Interactions Involving O3 and Plant Pathogens
o
o




1
O
H
6
o
0
H
O
o
0
o
H
W
Host Plant
Obligate Biotrophs
Bottle gourd (Lagenaria siceraria)
Cucumber (Cucumis saliva)
Pea (Pisum sativum)
Aspen (Populus tremuloides)
Hybrid poplar (Populus trichocarpa
x balsamifera)
Broad bean (Viciafaba)
Facultative Necrotrophs
Kidney bean (Phaseolus vulgaris)
Scots pine (Pinus sylvestris)


Pathogen

Powdery mildew
(Sphaerotheca fulginea)
Powdery mildew
(Sphaerotheca fulginea',
Powdery mildew
(Erysiphe polygoni)
Leaf rust (Melampsora
medusae f. sp.
tremuloidae)
Leaf rust (Melampsora
larici-populina or
M. allii-populina)
Bean rust (Uromyces
viciae-fabae)
Grey mold
(Botrytis cinerea)
Grey mold
(Botrytis cinerea)
White mold
(Sclerotinia sclerotiorum)
Annosus root and butt rot
(Heterobasidion annosum)


Effect of O3 on Disease

Increased in 50ppb O3;
decreased in 100+ppb
Increased in 50ppb O3;
decreased in 100+ppb
Decreased infection
Increased severity
Increased infection and
severity
Not reported
Increased from conidia on O3-
sensitive cultivar; decreased
from mycelium
Increased infection
Increased infection
Increased development*


Effect of Disease on O3 Response

Decreased; partial protection
Synergistic increase in 50ppb O3;
antagonistic decrease in 100+ppb;
partial protection
Decreased; partial protection
Not reported
Increased sensitivity (synergistic)
Decreased; partial protection
Not reported
Not reported
Not reported
Not reported


Reference

Khan and
Khan(1998a)
Khan and
Khan (1999)
Rusch and
Laurence
(1993)
Karnosky
et al. (2002)
Beare et al.
(1999a)
Lorenzini
et al. (1994)
Tonneijck
(1994)
Tonneijck and
Leone (1993)
Tonneijck and
Leone (1993)
Bonello et al.
(1993)



-------
3
Table 9-11 (cont'd). Interactions Involving O3 and Plant Pathogens
to
o
o



VO
oo

DRAFT-DO >
0
H
O
Host Plant
Facultative Necrotrophs (cont'd)
Loblolly pine (Pinus taeda)
Hybrid poplar (Populus deltoides x
nigra)
Hybrid poplar (Populus trichocarpa
x balsamifera)
Wheat (Triticum aestivum)

Nematodes
Tomato (Lycopersicon esculentum)
* Increase completely countered by


Pathogen

Pitch canker
(Fusarium subglutinans)
Canker (Septoria musiva
[=Mycosphaerella
populinum})
Leaf spot (Marssonina
tremulae)
Blotch (Septoria
nodorum)
Tan spot (Pyrenophora
tritici-repentis)

Root-knot nematode
(Meloidogyne incognita)
Effect of O3 on Disease Effect of Disease on O3 Response

Increased development Increased sensitivity
Increased incidence Not reported
Increased spore germination Not reported
and lesion growth after
lOOppb O3 (30 days);
decreased germination after
200ppb (15 days)
Increased infection Not reported
Increased infection of Not reported
disease-susceptible genotypes

Increased development Increased foliar injury; reduced plant
growth (synergistic)
Reference

Carey and
Kelley (1994)
Woodbury
et al. (1994)
Beare et al.
(1999b)
Tiedemann
and Firsching
(1993)
Sah et al.
(1993)

Khan and
Khan (1997,
1998a)
mycorrhizae (Hebeloma crustuliniforme).






O
HH
H
W

-------
 1      with numerous earlier reports), but only at high O3 exposures in the case of Sphaerotheca on
 2      cucumber. At low exposure levels, the disease and O3 acted synergistically.  The only other
 3      recent observation of such disease-related synergisms are the nematode-tomato reports of (Khan
 4      and Khan, 1997, 1998b).
 5           It is therefore clear that the type and magnitude of exposure to O3 plays an important role
 6      in determining both the response of the disease organism and of the host.
 7           No recent studies involving interactions between O3 and bacterial diseases appear to have
 8      been reported. With regard to viruses, a laboratory study by Yalpani et al. (1994) added to
 9      several reports of O3 decreasing the severity of tobacco mosaic virus  infection of tobacco,
10      Nicotiana tabacum; and Jimenez et al. (2001) reported that previous exposure to O3 resulted in
11      increased adverse effects on tomato yield attributed to several virus diseases.
12           Similarities between the sensitivities of different cultivars or clones to O3 and to specific
13      diseases have been  noted. For example, Sah et al. (1993) found that the severity of injury caused
14      by tar spot and standard O3 exposures of 12 wheat cultivars were closely correlated (R2 = 0.986).
15      Such similarities appear to have a mechanistic basis, since several studies have noted similarities
16      in the molecular and biochemical changes that occur in plants infected with pathogens and plants
17      exposed to O3. Schraudner et al. (1992), Ernst et al. (1992), Eckey-Kaltenbach et al. (1994a,b),
18      Yalpani et al. (1994) and Bahl et al. (1995) have presented evidence that exposures to O3 result
19      in responses such as increased levels of salicylic acid, the signaling agent for increased induced
20      resistance to pathogens.  This, in turn, leads to activation of the genes that encode defense
21      proteins, including the so-called pathogenesis-related proteins. The induction of such proteins
22      might account for the decreased infection with Sphaerotheca and Melampsora at higher O3
23      exposures but does  not account for increased infections seen at lower exposure levels. The issue
24      is discussed more fully by Sandermann (1996) and Schraudner et al. (1996).  More recently
25      Sandermann has extended the theory relating O3 and disease by suggesting that, because of O3
26      "memory effects" in affected host plants that may persist over weeks or months, analysis for
27      various induced biomarkers of gene activation may provide a useful tool for improving our
28      ability to predict the outcome of O3-plant-pathogen interactions (Sandermann, Jr., 2000).
29           There have been no reports of O3  studies with mixed infections by pathogens, but the
30      complete suppression of Heterobasidion butt and root rot of Scots pine by the mycorrhizal
        January 2005                              9-99        DRAFT-DO NOT QUOTE OR CITE

-------
 1      symbiont Hebeloma crustuliniforme indicates the possibility of interactions involving more than
 2      one fungus (see Section 9.4.4.3.3 below).
 3           In summary, our understanding of oxidant-plant-disease interactions is far from complete.
 4      However, a combined tabulation of the evidence presented in the 1996 O3 AQCD (U.S.
 5      Environmental Protection Agency, 1996) and that noted in Table 9-11 leads to the following
 6      summary of O3 effects on plant diseases and corollary effects of infection on plant response
 7      to O3, as indicated by number of studies showing increases or decreases in disease or
 8      susceptibility:
 9           For obligate biotrophic fungi, bacteria, nematodes:
10                O3 increased disease:   9          Increased susceptibility to O3:    3.
11                O3 decreased disease:   15         Decreased susceptibility to O3:    9.
12           F or facultative necrotrophi c fungi:
13                O3 increased disease:   25         Increased susceptibility to O3:    2.
14                O3 decreased disease:   3          Decreased susceptibility to O3:    4.

15      Thus, although O3 may reduce the severity but not the incidence of some of the diseases caused
16      by the obligate pathogens, the evidence overall indicates that with most diseases, their severity is
17      more likely to be increased by  O3 than not. However, the actual consequences will be specific to
18      the disease and level of exposure, and, most importantly, will be determined by environmental
19      suitability  and epidemiological requirements for disease to develop. Conversely,  some evidence
20      suggests that infection by obligate pathogens may confer some degree of "protection" against O3,
21      a dubious benefit from the plant's point of view.
22
23      9.4.3.3  Oxidant-Plant-Symbiont Interactions
24           No further studies have appeared regarding O3 effects on the important bacterial symbiont
25      of legumes, Rhizobium,  since those summarized in the 1996 document (U.S. Environmental
26      Protection Agency, 1996).  Hence, our present understanding is that, although relatively high
27      levels of exposure (>  200 ppb) can result in severe (> 40%) reductions in nodulation (and
28      therefore nitrogen-fixation) on soybean roots, lesser reductions in nitrogen-fixing may be caused
29      by lower exposures. However, the data are inadequate to attempt to define any quantitative
30      exposure-response relationships.

        January 2005                              9-100       DRAFT-DO NOT QUOTE OR CITE

-------
 1           There have been a few recent reports on O3-plant-mycorrhizae interrelationships. These
 2      have mostly involved seedlings of coniferous tree species.  A transient O3-induced stimulation of
 3      mycorrhiza on Scots pine roots reported by Kasurinen et al. (1999) was not observed in a later
 4      study by Kainulainen et al. (2000b).  Studies of the mycorrhiza Paxillus involutus on birch
 5      (Betulapenduld) seedlings showed that, although O3 reduced mycorrhizal growth rate, it led to
 6      greater extension growth which in turn resulted in greater mycorrhizal infection of neighboring
 7      Aleppo pine (Pinus halepensis) seedlings (Kytoviita et al.,  1999). However, O3 reduced nitrogen
 8      acquisition by P. halepensis from its mycorrhizal symbiont (Kytoviita et al., 2001). The
 9      complex interrelationships that may occur in the rhizosphere were revealed by the observation
10      (Bonello et al., 1993) that the mycorrhiza Hebeloma crustuliniforme could overcome the
11      O3-stimulated severity of root rot on Scots pine (Pinus sylvestris) caused by the fungus
12      Heterobasidion annosum (noted in Section 9.4.4.3.2).
13           In summary, the available evidence is far too fragmented and contradictory to permit
14      drawing any general conclusions about mycorrhizal impacts.  The negative effects of O3 on
15      mycorrhizae and their functioning that have been reported have not necessarily been found to
16      lead to deleterious effects on the growth of host plants.  Thus, little has changed from 1991 when
17      Dighton and Jansen asked: " Atmospheric Pollutants and Ectomycorrhizae: More Questions
18      than Answers?" (Dighton and Jansen, 1991). Because of their important roles in ecosystems,
19      mycorrhizae are further discussed in Section 9.5 below.
20
21      9.4.3.4  Oxidant-Plant-Plant Interactions: Competition
22           Plant competition involves the ability of individual plants to acquire the environmental
23      resources needed for growth and development: light, water, nutrients and space. Intraspecific
24      competition involves individuals of the same species, typically in monocultural crop situations,
25      while interspecific competition refers to the interference exerted by individuals of different
26      species on each other when they are in a mixed culture.
27           In cropping situations, optimal cultural practices for row spacing and plant density/row
28      tend to balance the negative effects of intraspecific competition and the goal of maximum yield.
29      Although interspecific competition is agriculturally undesirable when it involves weak
30      infestations, the use of mixed plantings may be agriculturally deliberate, e.g., grass-legume
        January 2005                             9-101        DRAFT-DO NOT QUOTE OR CITE

-------
 1      mixtures used for pasture or forage. In natural plant communities, monocultures are rare, and
 2      complex interspecific competition is the norm.
 3           Although weak competition is the largest global cause of crop losses, little is known about
 4      the impact of O3 on crop-weed interactions. The topic does not appear to have been investigated
 5      in recent years. We can only speculate as to the possible consequences of O3 exposure on weed
 6      competition based on our limited understanding of the effects on a few, mostly two-component
 7      mixtures of cultivated species.
 8           The tendency for O3-exposure to shift the biomass of grass-legume mixtures in favor of
 9      grass species, reported in the 1996 O3 AQCD (U.S. Environmental  Protection Agency, 1996) has
10      been confirmed by recent studies. In a ryegrass (Lolium perenne) + clover (Trifolium repens)
11      mixture grown in an open-air fumigation system, clover growth was impaired by extended
12      exposures to above-ambient O3,  leaving patches for weed invasion  (Wilbourn et al., 1995).
13      An open-top chamber study by Nussbaum et al. (1995b) using the same species confirmed the
14      greater effect on clover but observed that the magnitude of the effect depended highly on the
15      pattern of O3-exposures over extended growing periods.  Low-level exposures shifted species
16      composition in favor of Lolium.,  but exposures to higher peak O3  levels depressed total mixture
17      yield. With an alfalfa (Medicago saliva)  + timothy (Phleumprateme) mixture, Johnson et al.
18      (1996a) noted that O3 caused decreased alfalfa root growth and increased timothy shoot growth
19      and height. Nussbaum et al. (2000a) reported that, with increased exposure to O3, well-watered
20      red clover (Trifolium prateme) plants suffered from increased competition from the grass,
21      Trisetumflavescens, but the O3 exposure  also negatively effected grass growth, depressing
22      overall total yield. However, a greater adverse effect on Trisetum resulted from O3-induced
23      increased competition when grown with brown knapweed (Centaureajacea)., a weed species.
24           Andersen et al. (2001) demonstrated the potential for competition and O3-exposure to work
25      together to affect the growth of tree seedlings.  Ozone had no direct adverse effect on pine
26      growth in a 3-year study of ponderosa pine (Pinusponderosa} seedlings grown in mesocosms
27      with three densities of blue wild-rye grass (Elymus glaucus), but the O3-increased competitive
28      pressure of the grass caused a major reduction in pine growth.
29           Three studies have been reported on more complex plant associations. Ashmore and
30      Ainsworth (1995) studied mixed plantings of two grasses, Agrostis capilaris and Festuca rubra,
        January 2005                             9-102        DRAFT-DO NOT QUOTE OR CITE

-------
 1      with two forbs16, Trifolium repens (a legume) and Veronica chamaedrys, exposed to O3 in
 2      open-top chambers (OTCs).  The proportion of forbs, Trifolium in particular, declined, especially
 3      when cut at biweekly intervals.  In a related study, Ashmore et al. (1995) used artificial mixtures
 4      of grasses and forbs and transplanted swards of native calcareous grassland species and found
 5      that, regardless of whether total biomass was adversely affected by exposures to O3, higher
 6      exposures progressively shifted species composition, usually at the expense of the forb species.
 7      The observed shifts in competitive balance in favor of grasses is consistent with observations
 8      that many grass species are less sensitive to O3 than forbs.  However, as previously shown by
 9      Evans and Ashmore (1992), knowledge of the relative sensitivities to O3 of the component
10      species grown in isolation or in monoculture does not always predict the impact of O3 on the
11      components in  a mixed culture.
12          Barbo et al. (1998) exposed an early successional plant community to O3 in OTCs for two
13      growing seasons. Ozone decreased community structure features such as height of canopy,
14      vertical canopy density (layers of foliage), and species diversity and evenness. Surprisingly,
15      blackberry (Rubus cuneifolius), a species considered to be O3-sensitive, replaced sumac (Rhus
16      copallind) canopy dominance. Barbo et al. (2002) also demonstrated the role of competition in
17      determining the impact of O3 on loblolly pine (Pinus taedd).  They reported that the increased
18      growth of natural competitors in OTCs using charcoal-treated air to reduce the ambient O3
19      concentrations resulted in decreased pine growth. They noted that this is contrary to the
20      frequently reported increased growth observed in reduced O3 levels in the absence of
21      interspecific competition.
22          Competitively disadvantaged trees of four clones of aspen (Populus tremuloides) exposed
23      to 1.5* ambient O3 in a FACE facility over 4 years were proportionately more adversely affected
24      by O3 than competitively advantaged or neutral trees (McDonald et al., 2002). However, one
25      clone of the disadvantaged trees demonstrated enhanced growth.
26          In summary, our present knowledge of how O3 may affect the competitive interspecific
27      plant-plant relationships typifying the agricultural and natural worlds is very limited. However,
28      as noted in the  1996 AQCD (U.S.  Environmental Protection Agency, 1996), "the development
29      and use of field exposure systems  have permitted many recent studies of crop species to be
30      conducted at normal planting densities and hence have incorporated intraspecific competition as
              16Forb: any non-grassy herbaceous species on which animals feed.

        January 2005                             9-103       DRAFT-DO NOT QUOTE OR CITE

-------
 1      an environmental factor." Such facilities were used in most of the studies of interspecific
 2      competition discussed above. But we are still a long way from being able to bridge the gap
 3      between small model competing systems and the realities of natural ecosystem complexity.
 4
 5      9.4.4  Physical Factors
 6           The physical features of a plant's aerial and edaphic environments exercise numerous
 7      controls over its growth and development.  Thus, many of their effects may be modified by
 8      exposure to atmospheric oxidants and, alternatively, plants may modify responses to such
 9      exposures. As in the 1996 O3 AQCD (U.S. Environmental Protection Agency, 1996), this
10      section focuses on the defining features of plant microclimate: light,  temperature, relative
11      humidity (HR, or saturation vapor pressure deficit), and the presence  and availability of water,
12      especially in the soil. Monteith and Elston (1993) suggested that light energy and mass of water
13      should be viewed as climatic resources, and the other two elements (temperature and saturation
14      vapor pressure deficit) as rate modifiers that determine how fast the resources are used.  The
15      modifications of plant response by physical environmental factors has recently been reviewed by
16      Mills (2002).
17           Another physical feature of the microclimate, wind and air turbulence, which affects the
18      thicknesses of the boundary layers over leaves and canopies and, hence, affects gas  exchange
19      rates (including the fluxes of O3 and other  oxidants into the leaves) is discussed elsewhere
20      (Section 9.3).
21           Physical features of the environment are also important components of larger-scale
22      regional and global climates.  However, the following discussions are confined to issues related
23      to individual factors at the plant level; meso-scale  effects are reviewed in Section 9.4.4.8, which
24      addresses the issues of climate change interactions.
25
26      9.4.4.1  Light
27           Plants are the primary producers of biomass  on the planet through their ability to capture
28      light energy (by the process of photosynthesis) and convert it to the many forms of chemical
29      energy that sustain their own growth and that of secondary consumers and decomposers.  Light
30      intensity is critical since the availability of light energy (a resource, sensu [Monteith and Elston,
31      1993]) governs the  rate at which photosynthesis can occur, while light duration (i.e.,

        January 2005                              9-104        DRAFT-DO NOT QUOTE OR CITE

-------
 1      photoperiod) profoundly effects development in many species.  Although light quality (i.e., the
 2      distribution of incident wavelengths) may also affect some physiological plants processes, there
 3      is no evidence to indicate that such effects are of relevance to concerns over oxidant pollution,
 4      except at the short wavelengths of UV-B. This topic is discussed in the context of climate
 5      change in Section 9.4.8, and as  a stress factor per se affected by atmospheric O3 in Chapter 8.
 6      However, as noted above and in the 1996 O3 AQCD (U.S. Environmental Protection Agency,
 7      1996), none of the features is controllable in natural field situations. A brief discussion of light
 8      intensity-O3 interactions is included in the review by Chappelka and Samuelson (1998).
 9           The conclusion in the 1996 O3 AQCD (U.S. Environmental Protection Agency, 1996) that
10      low light intensities and short photoperiods tended to  increase susceptibility to foliar O3-injury
11      may still be valid, but this may  or may not translate into adverse effects on growth.  For
12      example, Tjoelker et al. (1993)  found that, when seedlings of sugar maple (Acer sacchamm),
13      a shade-tolerant species, were grown in 7% full  sunlight, O3 reduced shoot and root growth, but
14      had no significant effect in 45% sunlight (a 6-fold increase). In contrast, the reverse was
15      observed with a shade-intolerant hybrid poplar (Populus tristis x balsamifera\ with the greater
16      impact of O3 occurring in the higher light intensity treatment.
17           The greater sensitivity of maple in low light has also been confirmed in other studies.
18      Tjoelker et al. (1995) noted a greater O3-induced inhibition of photosynthetic CO2 assimilation in
19      shaded leaves than in leaves in full sunlight.  However, in the absence of differences in stomatal
20      conductance, the effect was considered to be independent of O3 flux; it appeared to be a
21      consequence of reduced chlorophyll contents and quantum efficiencies induced by O3.
22      In contrast, Back et al. (1999), who also observed a greater inhibition of net photosynthesis by
23      O3 in shaded leaves, reported decreased stomatal conductance. Although reduced conductance
24      might suggest reduced O3 flux and, therefore, decreased adverse effects, the authors concluded
25      that the effects of reduced conductance was offset by long-term changes in leaf structure, leading
26      to less densely packed mesophyll cells and greater internal air space within the leaves.
27      Morphological differences between lower and upper crown leaves of black cherry (Prunus
28      serotina) have been suggested as the basis for the greater O3-susceptibility of the  lower crown
29      leaves (Fredericksen et al., 1995). Back et al. (1999) also observed  accelerated foliar senescence
30      induced by O3 on shaded leaves, a response also noted by Topa et al. (2001). Sensitivity to O3
31      was found to be increased in shade- but not sun-leaves of shade-tolerant red oak (Quercus rubrd)

        January 2005                             9-105       DRAFT-DO NOT QUOTE OR CITE

-------
 1      (Samuelson and Edwards, 1993).  Similarly, Mortensen (1999) observed that seedlings of
 2      mountain birch (Betulapubescens) grown in 50% shade suffered greater foliar injury from O3
 3      than those grown in full sunlight.
 4           Not all shade-intolerant species exhibit greater reductions in photosynthesis and growth
 5      due to O3 when grown in full sunlight. Higher than ambient levels of O3 failed to inhibit
 6      photosynthesis in leaves of shade-intolerant yellow poplar (Liriodendron tulipifera) grown in
 7      nearly full sunlight (Tjoelker and Luxmoore, 1991). Greater foliar injury in the lower, shaded
 8      leaves of shade-intolerant black cherry (Primus serotind) trees and saplings, was attributed
 9      (Fredericksen et al., 1996a) to higher stomatal conductance and greater O3 uptake relative to net
10      photosynthetic rate. However, in a 3-year study of Norway spruce seedlings (Picea abies) in
11      OTCs.  Wallin et al. (1992) observed that photosynthetic efficiency was more adversely affected
12      by O3 in high than in low light.
13           The suggestion of greater sensitivity to O3  of shade-tolerant species in low-light conditions
14      and the greater sensitivity of shade-intolerant species in high light is somewhat of an over-
15      simplification when dealing with mature trees, for which light intensity varies considerably
16      within the canopy because of shading. Chappelka and Samuelson (1998) noted that the
17      interaction between sensitivity to O3 and the light environment in forest trees is further
18      complicated by developmental stage, with seedlings, saplings, and mature trees frequently giving
19      different results.  Topa et al. (2001) also cautioned that O3 effects on leaf-level photosynthesis
20      may be poor predictors of the growth responses of sugar maple in different light environments.
21           In high-light intensities, many species exhibit some degree  of photoinhibition of the
22      photosynthetic process through the overloading of the mechanisms that protect the
23      photosynthetic reaction centers in the chloroplasts.  Guidi et al. (2000) reported complex
24      interactions between high-light intensities (inducing photoinhibition) and O3 exposures in kidney
25      bean (Phaseolus vulgaris) with high intensities tending to enhance the detrimental effect of O3
26      on photosynthesis. One of the studies in the extensive European  Stress Physiology and Climate
27      Experiment-wheat (ESPACE-wheat) program (Bender et al., 1999), conducted in 1994 and
28      1995, included an investigation of the effects of climatic variables on yield response to O3 using
29      two simulation models, AFRCWHEAT2-O3 and LINTULCC (Ewart et al., 1999;  Van Oijen and
30      Ewart,  1999). Among the observed trends, it was noted that relative yield loss of wheat due to
31      elevated O3 tended to increase with light intensity.  In contrast, Balls et al. (1996) used ANNS to

        January 2005                             9-106       DRAFT-DO NOT QUOTE OR CITE

-------
 1      investigate microclimatic influences on injury caused by O3 to clover (Trifolium subterraneimi)
 2      and found that, especially at mid-range cumulative O3 exposures (350 to 500 ppb-h), injury
 3      tended to decrease with increasing light intensity. Similar observations by Davison et al. (2003)
 4      of foliar injury to wild populations of cutleaf cone flower (Rudbeckia laciniatd) exhibiting a
 5      range of PAR levels within their canopies led the authors to conclude that the variation in injury
 6      symptoms observed was "unlikely to be due to differences in ozone flux and more likely to be
 7      due to variation in light." Antonielli et al. (1997) found evidence indicating that the high
 8      sensitivity of the bioindicator tobacco cultivar Nicotiana tabacum cv Bel-W3 is partly
 9      determined by its high photosynthetic electron transport rates at high-light intensities, which
10      exceed the capabilities of the plant to dissipate energy and oxyradicals.
11           The  1996 O3 AQCD referred to the important role of light in controlling stomatal opening
12      and suggested that light duration (i.e., photoperiod) might dictate the actual uptake of O3 to some
13      degree.  However, it should also be noted that Sild et al. (1999) found that clover plants could
14      suffer foliar injury even if they were exposed to O3 during the dark period of the day-night cycle,
15      when stomatal conductance is at its lowest.
16           A possible indirect effect of light intensity was noted by Reiling and Davison (1992) in
17      their study of the O3-tolerance ofPlantago major plants grown from seeds collected from
18      populations at 28 different sites in Britain.  Ozone-tolerance, defined in terms of plant growth,
19      was found to be a function of both previous O3-exposure history  and hours of bright sunshine
20      during the year before the seeds were collected. However, the authors cautioned that, since
21      tropospheric O3-formation is itself dependent upon irradiation, the observation  does not
22      necessarily imply a direct effect of light intensity  on the plants' response to O3.
23           The only recent studies concerning interactions with light quality appear to be those
24      involving O3 and UV-B as a component of climate change. These are dealt with in Section
25      9.4.4.8.2.  The effects of photoperiod on response to O3 or the converse do not appear to have
26      received any recent attention.
27           Although the intensity, quality, and duration of light are not controllable in the natural
28      world, the interactions of O3 with light intensity, in particular, clearly have relevance to the
29      growth of shade-tolerant and shade-intolerant species in mixed forest stands. It appears that the
30      nature of light intensity-O3 interactions may depend upon the type of light environment to which
31      the species are best adapted, with increased light intensity increasing the sensitivity of light-

        January  2005                              9-107       DRAFT-DO NOT QUOTE  OR CITE

-------
 1      tolerant and decreasing the sensitivity of shade-tolerant species to O3.  Although there is
 2      certainly some evidence to the contrary, this hypothesis is a reasonable summation of current
 3      understanding with regard to O3-light intensity interactions.
 4
 5      9.4.4.2  Temperature
 6           "Temperature determines the start and finish and rate and duration of organ growth and
 7      development" (Lawlor, 1998). Such processes depend on fundamental physiological activities
 8      that are mostly enzyme-mediated and whose kinetics are directly affected by temperature. Since
 9      the processes of enzyme deactivation and protein denaturation also increase as temperatures rise,
10      each enzymatic process has a unique optimum temperature range for maximal function.
11      However, the optima for different processes within the plant vary appreciably and, hence, the
12      optimum temperature range for overall plant growth is one within which all of the individual
13      reactions and vital processes are collectively functioning optimally, not necessarily maximally.
14      Furthermore, individual features of plant development (e.g., shoot and root growth, flowering,
15      pollen tube growth, fruit set, seed development) have different specific optima, so that
16      differential responses to temperature occur, leading to temperature-induced developmental
17      changes. For example, despite increased assimilation, increased temperatures may result in
18      decreased grain yields of crops such as wheat, because the growing season is effectively
19      shortened by a more rapid onset of senescence (Van Oijen and Ewart,  1999).
20           Rowland-Bamford (2000) noted that a plant's response to temperature changes will depend
21      upon whether it is growing at its near optimum temperature for growth or its near maximum
22      temperature, and whether any increase in mean temperature results in temperatures rising above
23      the threshold for beneficial responses.  Impairment by O3 of any process may be thought of as
24      being analogous to a downward shift below and away from the temperature optimum or an
25      upward shift above and away from the optimum. Since a temperature rise toward the optimum
26      would result in a rate increase, the combined effects of O3 and such an increase might neutralize
27      each other, while the effects of O3 and a decrease in temperature would likely be additively
28      negative. Above  the optimum temperature, the situations would be reversed with the effects of
29      increased temperatures and O3 being additively negative, and decreasing temperatures
30      counteracting any negative effect of O3. Thus, it is difficult to generalize about the interactions
31      of temperature and O3 on overall plant responses such as growth in which the different

        January 2005                              9-108        DRAFT-DO NOT QUOTE OR CITE

-------
 1      temperature-rate relationships of different growth components are merged, because they depend
 2      upon the relationship of any temperature changes to the optimum for a species.
 3           Studies of the effects of temperature on the impact of O3 have increased recently because
 4      of an increased need to understand the consequences of global warming as a component of
 5      climate change. Direct interactions of temperature with O3 are reviewed here, but the issues are
 6      addressed again in Section 9.4.8.1 in relation to changes in atmospheric CO2 levels.
 7           The 1996 criteria document (U.S. Environmental Protection Agency, 1996) stressed the
 8      interdependence of the temperature within the tissues of the leaf (where the various temperature-
 9      sensitive processes occur) on three distinct components:  the ambient air temperature, the heating
10      effect of incident infrared radiation during the photoperiod, and the evaporative cooling effect
11      caused by transpirational loss of water. It also cautioned that, especially in experiments using
12      controlled environment chambers, the effects of temperature could well  be confounded with
13      those of humidity /vapor pressure deficit (VPD).  Temperature and VPD are strongly interrelated,
14      and VPD plays an important role in regulating stomatal transpiration. Because of the role that
15      evaporative cooling plays in determining internal leaf temperatures, any factor that causes
16      stomatal closure and reduced conductance inevitably leads to increased leaf temperatures. Such
17      interactions add to the difficulties in distinguishing the effects of temperature from those of other
18      factors, as actual  leaf temperatures are rarely measured and reported.
19           Despite these caveats, there is some evidence that temperature per se influences plant
20      response to O3. For example, in rapid-cycling Brassica (Brassica rapd) and radish (Raphanus
21      sativus), marked O3-inhibitions of growth were observed at low root temperatures (13 °C) but
22      not at 18°C (Kleier et al., 1998, 2001). With regard to air temperature, this was included in the
23      range of micrometeorological variables studied in several recent extensive field studies  and was
24      found to have a significant effect on response to O3 in most cases. Ball et al. (1996) used ANNs
25      in an analysis of the growth of clover (Trifolium subterraneuni) and concluded that light and
26      VPD had greater influences than temperature on the visible injury response to O3. However, in
27      three studies with different cultivars of white clover (Trifolium repens),  temperature was found
28      to be important to the growth response. Ball et al. (1998) exposed T. repens cv. Menna to
29      ambient O3 in OTCs at 12 European sites at a range of latitudes and altitudes from 1994 to 1996.
30      The impact of O3 on growth was determined as the ratio of growth with  and without treatment
31      with the O3-protectant, EDU (see Section 9.2). Artificial neural network analysis showed that O3

        January 2005                              9-109       DRAFT-DO NOT QUOTE  OR CITE

-------
 1      exposure (measured as the AOT40 index, see Section 9.4.6), VPD, and temperature were
 2      consistently the three most important variables governing response to O3 over a range of
 3      different ANN models. However, the authors did not describe the form of the O3-response
 4      relationship with temperature.  Similar observations were reported by Ball et al. (2000) and Mills
 5      et al. (2000) for O3-sensitive and O3-tolerant clones of T. repens cv. Regal, grown at 14 to
 6      18 European locations from 1995 to 1998.  In both studies, the impact of O3 was measured as the
 7      sensitive/tolerant growth (biomass) ratio. Although Ball et al. (2000) found temperature to be
 8      less important than O3 exposure and VPD, Mills et al. (2000) found temperature to be the most
 9      important input variable after O3 exposure (AOT40). In both cases, the adverse effect of O3
10      increased with increasing temperature.
11           A study of black cherry (Prunus serotind) seedlings and mature trees in Pennsylvania,
12      using micrometeorological variables aimed to predict O3 uptake, found temperature to be
13      unimportant (Fredericksen et al., 1996b), but in the study of populations ofPlantago major
14      referred to in Section 9.4.4.1, Reiling and Davison (1992) noted a weak, positive  correlation
15      between mean temperature at the collection site and O3 tolerance  (based on growth rate) of the
16      different populations.  In contrast,  Danielsson et al. (1999) collected genotypes ofPhleum
17      arvense from a wide range of Nordic locations and found a positive effect of temperature on the
18      growth of genotypes from locations wth higher summer temperatures, but sensitivity to O3 did
19      not vary systematically with geographic location.
20           Van Oijen and Ewart (1999) studied the effects of climatic variables on the  response to O3
21      in the ESP ACE-wheat program, based on two distinctive simulation models (AFRCWHEAT2-
22      O3 and LINTULCC, [Ewart et al.,  1999]) and noted that although the relative yield loss of wheat
23      due to elevated O3 tended to increase with temperature, the effect was of minor significance.
24           In contrast to the variable results obtained in studies of the effects of temperature on
25      response to O3, the corollary effect of O3 exposure on subsequent sensitivity to low temperature
26      stress, noted in the 1996 criteria document, is well recognized.  In reviewing low  temperature-O3
27      interactions, Colls and Unsworth (1992) noted that winter conditions produce three kinds of
28      stress: desiccation, chilling or freezing temperatures, and photooxidation of pigments. Of these,
29      they suggested that while the first two were important, the last may  play a particularly significant
30      role because the "combination of high irradiance and low temperatures permits a  buildup of free
31      radicals in leaf tissue, and these free radicals then attack chlorophyll."  Chappelka and Freer-

        January 2005                             9-110        DRAFT-DO NOT QUOTE OR CITE

-------
 1      Smith (1995) suggested that the injury and losses to trees caused by this delayed impact of O3
 2      may be equally or more important than the direct impacts of O3 on foliage of visible injury and
 3      necrosis, or the disruption of key physiological processes such as photosynthesis. In this
 4      context, the 1996 O3 AQCD (U.S. Environmental Protection Agency, 1996) referred to the
 5      conceptual framework (Eamus and Murray, 1991), which is still valid: brief periods of mild
 6      temperatures in the severest winters result in dehardening; O3 decreases frost hardiness per se,
 7      but also increases the predisposition to dehardening; dehardening places O3-exposed trees at
 8      greater risk from subsequent low temperatures. However, no quantified models of these effects
 9      have yet appeared.
10           The 1996 criteria document also noted that O3 adversely affects cold hardiness of
11      herbaceous species. More recently, Foot et al. (1996, 1997) observed winter injury and
12      decreased growth in low-growing perennial heather Calluna vulgaris exposed to O3 (70 ppb,
13      8 h/day, 5 days/week for 6 months) during the winter (6.8 °C mean), but found no significant
14      effects of the same exposures during the summer (12.3 °C mean).  Although Potter  et al. (1996)
15      observed a similar situation with the moss Polytrichum commune, the reverse was found with the
16      moss Sphagnum recurvum.
17           In summary, unequivocal evidence exists that O3 causes sensitization to the adverse effect
18      of low temperatures, but there is no clear pattern to evidence regarding the effects of temperature
19      on O3 response.  The many contradictory responses to temperature and O3 probably reflect our
20      lack of detailed knowledge of the temperature optima for the different growth components of the
21      studied species.  The topic of temperature-oxidant interactions is revisited later in Section 9.4.4.8
22      in the context of global warming as a feature of climate change.
23
24      9.4.4.3 Humidity and Surface Wetness
25           The moisture content of the ambient air (or its VPD) is a rate modifier (sensu  Monteith and
26      Elston, 1993) and an environmental regulator of stomatal conductance.  Both of the previous
27      criteria documents (U.S. Environmental Protection Agency, 1986, 1996) concluded that the
28      weight of evidence indicated that high RH (= low VPD) tended to increase the adverse effects of
29      O3, principally because the stomatal closure induced in most situations by O3 is inhibited by high
30      RH, leading to increased O3 flux into the leaves.
        January 2005                              9-111        DRAFT-DO NOT QUOTE OR CITE

-------
 1           Recent reports have confirmed this role of RH. The studies by Ball et al. (1995, 1996,
 2      1998) showed that VPD was an important determinant of O3-induced injury and reduced growth
 3      in two species of clover, Trifolium repens cv. Menna and T. subterraneum.  However, Mills et al.
 4      (2000) found it to be unimportant in the case of T. repens cv. Regal.  Such difference between
 5      cultivars is not unexpected since  considerable differences also occur among species and genera.
 6      For example, Bungener et al. (1999a) studied 26 Swiss grassland species and found clear
 7      evidence that O3 injury increased with decreased VPD (i.e., increased RH) in only eight species.
 8      However, the 1995 data from the European cooperative study of O3 injury, which involved
 9      28 sites in 15 countries and six crop species, led to the development of two 5-day critical-level
10      scenarios involving O3-exposure  (calculated as the AOT40 index) and mean VPD (0930-1630h):
11      200 ppb-h at > 1.5 kPa, and 500 ppb-h at < 0.6 kPa (Benton et al., 2000).
12           With forest tree species, Fredericksen et al. (1996b) found significant correlations between
13      stomatal conductance of black cherry (Prunus serotina) leaves and RH (+ve) and VPD (-ve),
14      and studies on free-standing Norway spruce (Picea abies) and larch (Larix decidud) showed that
15      although ambient VPD was highly positively correlated with ambient O3 concentration,
16      increased VPD caused stomatal closure, reducing O3 uptake and impact (Wieser and Havranek,
17      1993, 1995).
18           Surface wetness may affect the response to O3 through its direct effects on deposition to
19      the surface and through changes in RH.  Effects on the deposition of O3 have been reviewed
20      (Cape, 1996). A surface film of water on leaves was found to increase O3 deposition in four
21      studies involving field-grown grape (Vitis vinifera) (Grantz et al., 1995), red maple {Acer
22      rubruni) (Fuentes and Gillespie,  1992), deciduous forest dominated by largetooth aspen (Populus
23      grandidentata) and red maple (Fuentes et al., 1992), and clover-grass mixed pasture (Trifolium
24      pratense, Phleumpmtense, and Festucapratensis) (Pleijel et al.,  1995). In each case, the
25      increased deposition could be attributed partly to an increased stomatal conductance through the
26      abaxial (lower) surface and partly to uptake into the aqueous film on the adaxial (upper) surface.
27      In contrast, decreased deposition was noted by Grantz et al. (1997) with field-grown cotton
28      (Gossypium hirsutum). Since cotton is amphistomatous, with functional stomata on both leaf
29      surfaces, it was suggested that, in this case, the water layer effectively sealed the adaxial surface
30      stomata, more than offseting any increase in conductivity of the stomata in the abaxial surface.
31      However, none of the studies investigated the consequences of the differences in deposition.

        January 2005                             9-112       DRAFT-DO NOT QUOTE OR CITE

-------
 1      Although it could be inferred that, with part of any increased deposition being the result of
 2      increased O3 flux into the leaves, there would be the likelihood of increased O3 adverse effects,
 3      as suggested by earlier studies (Elkiey and Ormrod, 1981) that, by misting bluegrass (Poa
 4     pratemis) during exposure to O3, injury was significantly increased.
 5           To conclude, the effects of high RH (low VPD) and surface wetness have much in
 6      common, as they both tend to enhance the uptake of O3, largely through effects on stomata,
 7      leading to increased impact.
 8
 9      9.4.4.4  Drought and Salinity
10           The 1996 O3 AQCD (U.S. Environmental Protection Agency, 1996) concluded that the
11      available evidence clearly indicated that exposure to drought conditions could reduce the adverse
12      effects of O3 on the growth of herbaceous and woody plants, but it also noted that no quantitative
13      models of the O3-soil moisture deficit (SMD) interaction had yet appeared in print.
14      Nevertheless, the "protective" effect was inconsistent, and only appeared when SMD was
15      accompanied by high evaporative demand.  Since that time, further studies  have confirmed the
16      interaction, and simulation models have begun to appear.  Mills (2002) has recently provided a
17      brief review of the topic.
18           With regard to herbaceous species, Vozzo et al. (1995) observed less  O3-induced injury and
19      suppression of net photosynthesis and growth in water-deficient soybean (Glycine max) than in
20      well-watered plants. In several studies with wheat (Triticum aestivum), on  the other hand,
21      although adverse effects of both O3 and SMD were noted, they were consistently additive
22      (Bender et al., 1999; Fangmeier et al., 1994b, 1994a; Ommen et al., 1999).
23           In attempting  to model the stomatal conductance of wheat in relation  to O3 and soil
24      moisture, Griiters et al. (1995) found that although O3-induced stomatal closure was enhanced by
25      SMD, reducing O3 uptake, the R2 of the overall model was only 0.40, indicating that other
26      significant factors or relationships were involved.
27           With regard to native vegetation, Bungener et al. (1999a) used mixed plantings of 24 Swiss
28      grasses, herbs, and legumes and observed that, although O3-drought interactions were species-
29      specific, they tended to reflect stomatal functioning. They found that SMD reduced O3 injury in
30      two clovers (Trifolium repens and T. pratensis) and two grasses (Trisetum flavescens and
31      Bromus erectus\ but noted no interactions in the other 20 species. With relative growth rate as

        January 2005                             9-113       DRAFT-DO NOT QUOTE OR CITE

-------
 1      the measure of response to O3, interactions with SMD were noted in only three species:
 2      Trifolium repens and two weedy herbs, Knautia arvense and Plantago lanceolata (Bungener
 3      et al., 1999b). Although this variability in response among species was noted in the review by
 4      Davison and Barnes (1998), they also pointed out that in severely draughted regions of Europe,
 5      notably in Greece and Spain, O3-induced injury and growth reductions were common on many
 6      (usually irrigated) crops, but there were virtually no records of injury symptoms in wild species.
 7           Thus, the situation with herbaceous species is essentially unchanged from 1988 when
 8      Heagle et al. summarized the extensive NCLAN experiments that incorporated water stress as a
 9      variable:  "SMD can reduce the response of crops to O3 under some conditions but not under
10      other conditions. Probably the occurrence of O3 by SMD interactions was dependent on the
11      degree of SMD-induced plant moisture stress."
12           With regard to trees, O3 interactions with soil water availability have been discussed in
13      several recent reviews: Chappelka and Freer-Smith (1995), who focused on O3-induced
14      predisposition to drought stress; Johnson et al. (1996b); Chappelka and Samuelson (1998); and
15      Skarbyetal. (1998).
16           Several recent studies with conifers have yielded mixed results. No interactions with
17      drought were observed by Broadmeadow and Jackson (2000) on Scots pine  (Pinus sylvestris), by
18      Karlsson et al. (2002) on Norway spruce (Picea abies),  or by Pelloux et al. (2001) on Aleppo
19      pine (P. halepensis).  More recently, Le Thiec and Manninen (2003) reported that drought
20      reduced O3-induced growth suppression of Aleppo pine seedlings. Panek and Goldstein (2001)
21      inferred less impact of O3 on draughted Ponderosa pine (P. ponder osd), and Van den Driessche
22      et al. (1994) reported that drought reduced injury and O3-induced ethylene release by Picea
23      abies. But Karlsson et al. (1997), in a comparative study of fast- and slow-growing clones of
24      P. abies, only observed a drought-induced reduction of O3-inhibited root growth in the fast-
25      growing clone. In contrast, Grulke et al. (2002) noted a synergistic interaction between O3 and
26      drought stress on gross photosynthesis of Pinus ponder osa, and Wallin et al. (2002) reported a
27      synergistic growth response of Picea abies in the third year of a 4-year study. A similar
28      response was noted by Dixon et al. (1998) with the Istebna strain of P. abies.
29           With broad-leaved trees, studies of Durmast oak (Quercuspetraed) (Broadmeadow et al.,
30      1999; Broadmeadow and Jackson, 2000) and European ash (Fraxinus excelsior) (Broadmeadow
31      and Jackson, 2000; Reiner et al., 1996) showed that drought provided partial protection against

        January 2005                             9-114       DRAFT-DO NOT QUOTE OR CITE

-------
 1      O3-induced growth reduction. Although European beech (Fagus sylvaticd) is reportedly an
 2      O3-and drought-sensitive species, neither Pearson and Mansfield (1994) nor Broadmeadow et al.
 3      (1999) observed any interactions between these stresses, while Dixon et al. (1998) observed
 4      partial protection. Paakonen et al. (1998) observed only additive effects in a sensitive clone of
 5      birch (Betulapenduld).  However, the experiments of Schaub et al. (2003) and the survey by
 6      Vollenweider et al. (2003) on black cherry (Prunus serotind) cleary indicate antagonism between
 7      drought and O3 stresses on this species.
 8           With regard to the converse effect, in a critical review of the evidence for predisposition to
 9      drought stress being caused by O3, Maier-Maercker (1998) supported the hypothesis and
10      suggested that the effects were caused by the direct effects of O3 on the walls of the stomatal
11      guard and subsidiary cells in the leaf epidermis, leading to stomatal dysfunction.
12           The Plant Growth  Stress Model (PGSW) developed by Chen et al. (1994) is a physiology-
13      based process model which includes drought among several environmental variables.
14      Simulations for Ponderosa pine (Pinus ponder osd) incorporated antagonistic effects between O3
15      and drought stresses, i.e. partial protection, although Karlsson et al. (2000) have since
16      emphasized that drought-induced "memory effects" should be considered when developing
17      simulation models incorporating stomatal conductance.
18           Retzlaff et al.  (2000) used the single-tree model,  TREEGRO, to simulate the combined
19      effects of O3 and drought on white fir (Abies concolor). Although simulated reductions in
20      precipitation > 25% reduced growth, they also reduced O3 uptake (and impact). But lesser
21      reductions  in precipitation combined synergistically with O3 stress to reduce growth, leading the
22      authors to conclude that moderate drought may not ameliorate the response of white fir to O3.
23           On a much larger scale with a modified forest ecosystem model (PnEt-II) incorporating O3-
24      response relationships for hardwood species, Ollinger et al. (1997) showed how predicted
25      changes in net primary production and mean wood production in the northeastern U.S. hardwood
26      forests due to O3 would be reduced (but not countered  or reversed) by drought stress, particularly
27      in the southern part of the region. This geographic distribution of the effect was substantiated by
28      the work of Lefohn et al. (1997) on the risk to forest trees in the southern Appalachian
29      Mountains, based on localized estimates of O3 levels and SMD. The TREEGRO and ZELIG
30      models were combined by Laurence et al. (2001) to predict the impacts of O3 and moisture (as
31      precipitation) on the growth of loblolly pine (Pinus taedd) and yellow poplar (Liriodendron

        January 2005                             9-115        DRAFT-DO NOT QUOTE OR CITE

-------
 1      tulipiferd).  Based on O3 and precipitation data from three sites in the eastern United States, the
 2      six model regressions developed for the two species included both positive and negative
 3      coefficients for O3 exposure and precipitation as determinants of growth.
 4           As noted in the 1996 O3 AQCD (U.S. Environmental Protection Agency, 1996), the effects
 5      of soil salinity are similar to those of SMD. In a study of rice (Oryza sativci) cultivars of
 6      differing sensitivity to salinity, Welfare et al. (1996) noted that although both O3 and salinity
 7      reduced many features of growth additively, antagonistic interactions were only seen for leaf
 8      length and potassium accumulation.  Similarly, a recent study on chickpea (Cicer arietinum)
 9      found no interactions with regard to most components of biomass accumulation (the effects
10      of O3 and salinity were additive), but with root growth, salinity suppressed the adverse effects
11      ofO3.
12           In summary, the recently described interactions of O3 and drought/salinity stresses are
13      consistent with the view that, in many species, drought/salinity reduces the impact of O3 but O3
14      increases sensitivity to drought stress, i.e., the type of response is determined by the sequence of
15      stresses. However, synergisms have also been observed and any antagonisms are species-
16      specific and unpredictable in the absence of experimental evidence. In no case has an
17      antagonism been found to provide complete protection.
18
19      9.4.5  Nutritional Factors
20           The 1996 criteria document (U.S. Environmental Protection Agency, 1996) noted that the
21      large number of macro- and micronutrients and the wide range of species had prevented
22      experimental investigation of all but a few cases of nutrient-O3 interactions and most of these
23      concerned nitrogen (N) and crops or forest tree species. The document also provided a
24      comprehensive tabulation of the results of the relevant studies up to 1992.
25           The suboptimal supply of mineral nutrients to plants leads to various types of growth
26      reductions. The consequences of suboptimal nutrition might, therefore, be expected to have
27      some similarities to those of O3 exposure.  One might expect nutritional levels below the
28      optimum either to amplify any effects of O3 or at least lead to additive responses. The  difficulty
29      with this suggestion is that the available information has mostly been obtained from
30      experimentation conducted using two or more arbitrarily selected levels of fertility with little or
31      no regard to optima. Hence, it is not surprising that there have been contradictory reports, even

        January 2005                             9-116       DRAFT-DO NOT QUOTE OR CITE

-------
 1      among studies with the same species or cultivars conducted by different workers at different
 2      locations using different soils or soil mixes.
 3           There appear to have been no recent studies on O3 interactions with specific mineral
 4      nutrients other than N. Hence, the previous conclusions are still valid, viz. that increasing levels
 5      of the major elements potassium (K) and sulfur (S) usually reduce the impact of O3, or,
 6      deficiency increases susceptibility, whereas increased phosphorus (P) usually increases injury,
 7      or, deficiency decreases susceptibility.
 8           However, with N, a relationship to the optimum is usually demonstrable.  Several earlier
 9      studies of O3 x N interactions reported that the adverse effects of O3 on growth were greatest at
10      the optimum and decreased with increasing N-deficiency, a finding supported by the work on
11      aspen (Populus tremuloides) of Pell et al. (1995), who also confirmed that excess N decreased O3
12      impact on growth. Similarly, the adverse effects of O3 on growth rate in wheat (Triticum
13      aestivum) diminished with decreased N supply (Cardoso-Vilhena and Barnes, 2001).  However,
14      the effects of N are far from consistent. For example, Greitner et  al. (1994) reported that O3 and
15      N-deficiency acted additively in aspen in reducing leaf surface area and rate of photosynthesis,
16      Bielenberg et al. (2001) reported that the rate of O3-induced senescence was increased by
17      N-deficiency in hybrid poplar (Populus trichocarpa x maximovizii), and Paakkonen and
18      Holopainen (1995) observed the least adverse effects of O3 on birch (Betulapenduld) at
19      optimum N-fertility levels. With cotton (Gossypium hirsutum), increased N-levels  more than
20      overcame the adverse effect of O3 on growth and boll yield (Heagle et al., 1999). In view of
21      these contradictions, one may conclude that other, unrecorded factors may have contributed to
22      the various findings. Thus, much remains unclear about O3 x N-fertility interactions.
23           There have been two recent studies on the effects of overall  soil fertility.  Whitfield et al.
24      (1998) observed that low general fertility increased O3 sensitivity in selections ofPlantago
25      major. At the biochemical level, well-fertilized birch (Betulapenduld) saplings were found to
26      be less adversely affected by O3 than nutrient-stressed plants (Landolt et al., 1997).
27           TREEGRO model simulations of the growth of red spruce (Picea rubrd) in conditions of
28      nutrient  deficiency and O3 stress showed that, in combination, the two stresses acted less than
29      additively (Weinstein and Yanai, 1994). Minimal amelioration by nutrient deficiency was
30      predicted with Ponderosa pine (Pinusponderosa).
        January 2005                             9-117        DRAFT-DO NOT QUOTE OR CITE

-------
 1           Plants may also obtain N and S from airborne sources such as NOX, HNO3, NO3 , SO2, and
 2      SO42 , although, depending upon their concentration, these may also be phytotoxic.  In various
 3      parts of the world, the deposition of N and S in these forms contributes significantly to the levels
 4      of nutritionally available N and S in soils. Such depositions may, in turn, influence the impact of
 5      O3 on sensitive species through their roles as nutrients independent of any interactions that may
 6      occur because of their acidic properties (see Section 9.4.6.5). For example, Takemoto et al.
 7      (2001) recently reviewed the situation in southern California's mixed conifer forests and noted
 8      that, where N-deposition is appreciable, its  combination with O3 is causing a shift in Ponderosa
 9      pine (Pinusponderosd) biomass allocation towards that of deciduous trees, with increased
10      needle drop so that only 1- and 2-year needle classes overwinter.  Such changes are having
11      significant consequences to the balance of the forest ecosystem and are discussed more fully in
12      Section 9.5.
13           Of the micronutrient elements,  only manganese (Mn) appears to have been studied
14      recently.  In beans (Phaseolus vulgaris) Mn-deficiency increased O3 toxicity, despite causing
15      reduced O3 uptake (through decreased stomatal conductance) and inducing increased levels of
16      Mn-SOD (Mehlhorn and Wenzel, 1996).
17           In view of the foregoing, it is impossible to generalize about the interactions of soil fertility
18      with O3. While this is especially true of the interactions involving soil nitrogen, for which there
19      is much conflicting evidence, the interactions with other nutrients need much more thorough
20      investigation than has occurred to date, before any clear patterns become apparent.
21
22      9.4.6  Interactions  with Other Pollutants
23           The ambient air may be polluted by gases other than O3 and its photochemical  oxidant
24      relatives. In particular,  industrial, domestic, and automobile emissions and accidents can lead to
25      significant atmospheric concentrations of gases such as sulfur dioxide (SO2), nitric oxide (NO),
26      and nitrogen dioxide (NO2), collectively referred to as NOX, both locally and regionally. Local
27      releases of gases such as hydrogen fluoride (HF), hydrogen chloride (HC1), and chlorine (C12)
28      may result from industrial emissions and accidents.  Agricultural fertilizer and manure usage can
29      lead to significant increases in ambient ammonia (NH3). The sulfur and nitrogen oxides may
30      undergo reactions in the atmosphere leading to the formation of sulphate (SO42 ) and nitrate
31      (NO3 ) ions and resultant acid deposition.

        January 2005                             9-118        DRAFT-DO NOT QUOTE OR CITE

-------
 1           The 1996 O3 AQCD (U.S. Environmental Protection Agency, 1996) discounted much of
 2      the early research on pollutant combinations, because of its lack of resemblance to the ambient
 3      experience: the concentrations used were unrealistically high or the exposure regimes employed
 4      almost invariably used gas mixtures, whereas Lefohn et al. (1987) showed that the co-occurrence
 5      patterns of significant levels of O3 with SO2 or NO2 in the United States were most frequently
 6      sequential or partially sequential with overlap; only rarely were they entirely concurrent. On the
 7      other hand, O3 and peroxyacetylnitrate (PAN) frequently co-occur, as both form photochemically
 8      under similar conditions.
 9           To the list of reviews mentioned in the 1996 criteria document should be added the more
10      recent ones by Barnes and Wellburn (1998), Robinson et al. (1998), and Fangmeier et al. (2002),
11      which also explore some of the potential mechanisms underlying pollutant-pollutant interactions.
12
13      9.4.6.1   Oxidant Mixtures
14           In 1998, Barnes and Wellburn noted that virtually no information existed on the effects on
15      plants of concurrent exposures to O3 and other components of photochemical oxidant other than
16      PAN. The situation has not changed since their review appeared, and the topic appears to have
17      attracted no research interest since before the 1996 O3 AQCD (U.S. Environmental Protection
18      Agency, 1996). The continuing conclusion must, therefore, be that, from the limited information
19      available, the two gases appear to act antagonistically, with O3 raising the threshold for the
20      visible injury response to PAN, and PAN reducing the harmful effects of O3.
21
22      9.4.6.2   Sulfur Dioxide
23           In reviewing O3 x  SO2 interactions, Barnes and Wellburn (1998) remarked: "The outcome
24      of exposure to this combination of pollutants has probably been the most  studied, yet is one of
25      the least understood." More recent studies have only added to the conflicts referred to in the
26      1996 criteria document (U.S. Environmental Protection Agency, 1996), rather than resolve them.
27      For example, Diaz et al. (1996) reported that, after a year of daily exposures of Aleppo pine
28      (Pinus halepensis)sQQd\ings to 50 ppb O3 and/or 40 ppb SO2, the combination of pollutants
29      synergistically reduced shoot and root growth and impaired mycorrhizal colonization of the
30      roots. With tomato (Lycopersicon esculentum\ on the other hand, effects on growth ranged
31      from synergistic at low exposures (50 ppb) to antagonistic at exposures of 200 ppb of each gas

        January 2005                            9-119       DRAFT-DO NOT QUOTE OR CITE

-------
 1      (Khan and Khan, 1994). Although various physiological measurements were made in these and
 2      earlier studies, it has not been possible to determine any consistent mechanism or mechanisms
 3      that might account for the conflicting results.
 4           Since the information available about O3 x SO2 interactions appears to be highly dependent
 5      upon species, the type of response measured, and the experimental protocol used, it would still
 6      appear prudent to heed the statement of Heagle et al. (1988) in their summary of the studies
 7      undertaken in 12 field experiments over several years within the NCLAN program: "There were
 8      no cases where O3 and SO2 interactions significantly affected yield." (emphasis added.)
 9
10      9.4.6.3   Nitrogen Oxides, Nitric Acid Vapor, and Ammonia
11           The major oxides of nitrogen that occur in ambient air are nitrous oxide (N2O), nitric  oxide
12      (NO) and nitrogen dioxide (NO2), of which the latter two (conveniently symbolized as NOX) are
13      particularly important in connection with O3, because they are components of the reaction mix
14      that leads to photochemical O3 formation and because they can interact with O3-responses.  Their
15      reactions in the atmosphere can also lead to the occurrence of nitric acid vapor (HNO3) in
16      ambient air. The other major N-containing contaminant of ambient air in many parts of the
17      world is ammonia (NH3), largely released through agricultural practices.
18           Despite various combinations of O3 and NOX being probably the most common air
19      pollutant combinations found in the field, Barnes and Wellburn (1998) noted that they have been
20      little studied. Much early work with O3 and NOX focused on O3  x NO2 interactions and can be
21      discounted, because of the unrealistic concentrations employed and their use as mixtures rather
22      than in types of sequences.  The 1996 O3 AQCD (U.S. Environmental Protection Agency, 1996)
23      concluded that evidence from studies involving concurrent exposures to both O3 and NOX at
24      realistic concentrations was so fragmented and varied that no firm conclusions could be drawn as
25      to the likelihood and nature such interactions.  However, the few recent investigations taken
26      together with the earlier data are now beginning to reveal a pattern of response.
27           With regard to NO, Nussbaum et al. (1995a; 2000b) reported their findings with concurrent
28      exposures to NO and O3 and observed that, at low O3 levels, NO tended to act similarly to O3 by
29      increasing the scale of responses such as growth reductions. However, in ambient air in which
30      O3 is a dominant factor, the effects of NO were usually found to be negligible due to low levels,
        January 2005                              9-120       DRAFT-DO NOT QUOTE OR CITE

-------
 1      although the authors admitted that the effects observed were confounded by the inevitable
 2      O3-induced oxidation of NO to NO2.
 3           Two possible mechanisms whereby NO may influence plant response to O3 are suggested
 4      by recent biochemical studies.  First, there is growing evidence for the role of NO as a signaling
 5      agent in plants that can induce defense responses to a range of biotic and abiotic stressors
 6      (Beligni and Lamattina, 2001; Neill et al., 2002).  Second, a role for NO as an antioxidant
 7      scavenger of reactive oxygen species has been demonstrated by Beligni and Lamattina (2002) in
 8      potato (Solarium tuberosum) leaves and chloroplasts. However, both of these cases concern
 9      endogenously synthesized NO, and it must be noted that in none of these or other reports of
10      studies of NO signaling have the authors considered the potential significance of exogenous NO
11      in ambient air.
12           An independent case for O3 x NO interactions  comes from Mills et al. (2000). The ANN
13      model developed to predict the O3 effects on white clover (Trifolium repens) biomass based on
14      experiments at 18 locations throughout Europe found that the minimum daily NO concentration
15      (at 5 p.m.) was a significant contributor to adverse effects.
16           Turning to NO2, Maggs and Ashmore (1998) found that, although concurrent but
17      intermittent exposures of Bismati rice (Oryza saliva) revealed no significant growth interactions,
18      NO2 reduced the rate of O3-induced senescence, an antagonistic response possibly related to
19      enhanced N-metabolism.
20           With regard to sequential exposures, two studies on gene  activation in tobacco (Nicotiana
21      tabacum) revealed that NO2 counteracted the effect of O3 in reducing mRNA levels for three
22      genes encoding photosynthetic proteins (Bahl and Kahl,  1995) and tended  to counteract the
23      O3-induced enhancement of defense-protein gene activation (Bahl etal., 1995). However,
24      despite compelling evidence for significant interactive effects provided by earlier studies
25      (Bender et al., 1991; Goodyear and Ormrod, 1988; Runeckles and Palmer,  1987), the only recent
26      investigation of growth effects seems to have been that of Mazarura (1997) using sine-wave
27      exposure profiles.  He found that although 4 weeks of twice daily 3-h exposures to NO2 (120 ppb
28      peak concentrations) slightly stimulated growth of radish (Raphanus saliva] and while daily 6-h
29      exposures to O3 (120 ppb peak concentration) did not significantly reduce  growth, the daily
30      sequence, NO2 - O3 - NO2, led to a 13% drop in dry matter production.
        January 2005                              9-121        DRAFT-DO NOT QUOTE OR CITE

-------
 1           The combined evidence to date, therefore, suggests that, in leguminous species, the effects
 2      of these sequences are antagonistic with NO2 tending to reduce (or reverse) the negative effects
 3      of O3 on growth, while the effects are increased in other species. These conclusions differ from
 4      those of Barnes and Wellburn (1998) who suggested that sequential exposures tended to result in
 5      antagonistic effects (largely based on the  summary by Bender and Weigel, 1992), whereas
 6      simultaneous exposures were likely to lead to synergistic responses. With disagreements both
 7      among the data and their interpretation, it is not possible to determine the circumstances under
 8      which specific interactions of O3 and NO2 may occur, but there is no reason to doubt the validity
 9      of the individual findings of each study. Far more systematic investigation is needed to clarify
10      the situation.
11           There appear to have been no studies of O3 interacting with HNO3 in the vapor phase.
12      However, in the southern California montane forests (Takemoto et al., 2001), in Sweden (Janson
13      and Granat, 1999), and elsewhere, significant amounts of N are deposited in this form because of
14      the vapor's high deposition velocity.  As a consequence, although much of it ultimately reaches
15      the ground through leaching and leaf fall  and enters the soil as nitrate (NO3 ), it may also be used
16      as a N source by the foliage itself.  This nutritional role is independent of any contribution that
17      HNO3 vapor may make to acidic deposition.  Indirect interactions with the effects of O3 through
18      N-deposition of NOX, HNO3, and NH3 are related to the interactions of O3 with N as a nutrient,
19      and have recently been examined in the review by Takemoto et al.  (2001). The 1996 O3 AQCD
20      (U.S. Environmental Protection Agency,  1996) stated that the evidence available at that time led
21      to estimates of total forest dry deposition, including HNO3, ranging from 5.7 to 19.1 kg N ha"1
22      year1 (Taylor, Jr. et al., 1988).  However, Takemoto et al. (2001) pointed out that in parts of the
23      mid-elevation forests of southern California, dry deposition rates may reach more than 40 kg N
24      ha"1 year1. As a result, some locations have seen the conversion from N-limited to N-saturated
25      forests. The concern for California's forests is well-stated by Takemoto et al.: "As potential
26      modifiers of long-term forest health, O3 is a stressor and N deposition is an enhancer of
27      ponderosa/Jeffrey pine physiology and growth (Grulke and Balduman, 1999). The  progression
28      toward a deciduous growth habit, higher shootroot biomass ratios, increasing depths of litter,
29      tree densification, and elevated NO3- levels in soil and soil  solution, all point to the replacement
30      of pine species with nitrophilous, shade- and O3-tolerant tree species, such as fir and cedar
31      (Minnich et al., 1995) (Minnich, 1999)."

        January 2005                              9-122        DRAFT-DO NOT QUOTE OR CITE

-------
 1           Few studies have been reported of interactions of O3 with NH3.  The 1996 criteria
 2      document made reference to the work on bean (Phaseolus vulgaris) by Tonneijck and van Dijk
 3      (1994, 1998). Although NH3 alone tended to increase growth and O3 alone to inhibit it, one
 4      interaction was noted (Tonneijck and van Dijk, 1994) on the number of injured leaves. Dueck
 5      et al. (1998) studied the effects of O3 and NH3 on the growth and drought resistance of Scots
 6      pine (Pinus sylvestris). Significant interactions were found for some growth features, but there
 7      were no consistent patterns of the effects of NH3 on O3 response or vice versa. However, O3 was
 8      found to ameliorate the enhancement of drought stress caused by NH3.
 9           At this time there is insufficient information to offer any  general conclusions about the
10      interactive effects of O3 and NH3.
11
12      9.4.6.4  Hydrogen Fluoride and Other Gaseous Pollutants
13           Although HF and other fluorides are important local air pollutants associated with
14      aluminum smelting and superphosphate fertilizer manufacture, no studies of possible interactions
15      with oxidants appear to have been reported since that of MacLean (1990). He found that HF
16      retarded the accelerated senescence and loss of chlorophyll resulting from O3 exposure in corn
17      seedlings. However, such an isolated observation cannot be taken to indicate that HF can reduce
18      the impact of O3 on other species or even that the effect would ultimately have led to an effect on
19      mature plants.
20
21      9.4.6.5  Acid Deposition
22           The deposition of acidic species onto vegetation may elicit direct effects on the foliage or
23      indirect effects via changes induced in the soil. The 1996 O3 AQCD (U.S. Environmental
24      Protection Agency, 1996) included an extensive listing of investigations into the effects of O3
25      and acid deposition (usually in the form of simulated acid rain, SAR) on plant growth and
26      physiology. The majority of studies found no effects of SAR or acid mists or fogs at pH values
27      greater than about 3.0 and no interactive effects with O3.  (In ambient air, pH values less than 3.0
28      have rarely been reported.) In the few reports in which significant interactions were found, most
29      were antagonistic and were explained as probably being the result of increased fertility due to
30      nitrate and sulfate supplied in the SAR.
        January 2005                             9-123        DRAFT-DO NOT QUOTE OR CITE

-------
 1           Although numerous reviews have recently appeared (e.g., Bussotti and Ferretti, 1998;
 2      Fliickiger et al., 2002; Fox and Mickler, 1996; Nussbaum et al., 1999; and Sheppard and Cape,
 3      1999), the shift in interest in air pollution effects away from acid deposition has resulted in little
 4      new research having been reported over the past 10 or so years. In most of the reported studies,
 5      no effects due to the O3 exposures, the SAR treatments used, or their combinations were
 6      observed, e.g., Baker et al. (1994) on loblolly pine (Pinus taeda); Laurence et al. (1997), and
 7      Vann et al. (1995) on red spruce (Picea rubens); and Laurence et al. (1996) on sugar maple
 8      (Acer saccharum).  Branch chamber studies of 12-year old Ponderosa pine (Pinusponderosd)
 9      trees by Momen et al. (1997, 1999) revealed no O3 effects or interactions.  With red spruce
10      (Picea rubens), Sayre and Fahey (1999) noted no effects of O3 on the foliar leaching of Ca or
11      Mg, which only became significant with SAR at pH 3.1. Izuta (1998) observed no interactions
12      with Nikko fir (Abies homolepis),  although SAR at ph 4.0 reduced dry matter.  Shan et al. (1996)
13      reported adverse effects of O3 but  none attributable to SAR on the growth  of Pinus armandi.
14           With herbaceous species, Ashenden et al. (1995) noted significant antagonistic interactions
15      of O3 and acid mist in white clover (Trifolium repens\ in which the adverse effect of low pH was
16      countered by O3.  In contrast, Ashenden et al. (1996) found that, although pH 2.5 mist caused a
17      significant stimulation of the growth of ryegrass (Lolium perenne) attributed to a fertilizer effect,
18      and O3  caused reduced growth, there was no interaction. Bentgrass (Agrostis capillaris) behaved
19      similarly.
20           A study by Bosley et al. (1998) on the germination of spores of the moss, Polytrichum
21      commune, and the ferns, Athyrium felix-femina and Onoclea sensibilis, revealed no effect of O3
22      on moss spores, while SAR at pH  < 4.0 was completely  inhibitory. With the ferns, germination
23      was progressively reduced by both increased  O3 and acidity.
24           In summary, the few findings of interactions in these recent studies are consistent with the
25      previous conclusion regarding the likelihood  of such interactions being antagonistic.  However,
26      the interactions observed were in each case largely the result of the response to the lowest pH
27      used, which in several cases was below 3.0, and hence may not be relevant to most field
28      conditions.
29
30
        January 2005                              9-124       DRAFT-DO NOT QUOTE OR CITE

-------
 1      9.4.6.6  Heavy Metals
 2           Since there appears to have been no further research into the interactions of oxidants with
 3      heavy metal pollutants, our understanding is unchanged from at the time of the 1996 O3 AQCD
 4      (U.S. Environmental Protection Agency, 1996). As noted therein, the limited data available
 5      from early studies indicates varying degrees of enhancement of any adverse effects of O3 but
 6      precludes the development of any response relationships.
 7
 8      9.4.6.7  Mixtures of Ozone with Two or More Pollutants
 9           In many airsheds the mixtures that occur, both concurrently and over time, may involve
10      three or more pollutants.  Very little useful information exists  on the effects of O3 with multiple
11      pollutants. As the 1996 criteria document and others have pointed out, most of the early studies
12      on such combinations can be discounted because of their use of (1) high and environmentally
13      irrelevant exposure concentrations and (2) unrealistic, repetitive exposure profiles (U.S.
14      Environmental Protection Agency,  1996; Barnes and Wellburn, 1998).
15           The large investment in experimental facilities required to study these complex interactions
16      is a major deterrent.  So, although the topic has been included in several reviews that have
17      appeared in the last decade, there appear to have been only two studies that have provided new
18      information on the effects of O3 in combination with more than one other pollutant stress.
19      Ashenden et al. (1995, 1996) studied the effects of O3 and/or (SO2 + NO2) with four acidities of
20      SAR applied to each gas treatment, on white clover (Trifolium repens) and two pasture grasses
21      (Lolium perenne andAgrostis capillaris).  With each species, the antagonism reported for the O3
22      x SAR interaction (Section 9.4.4.6.5) tended to be nullified by concurrent exposure to the other
23      gases, while the combination of the three gaseous pollutants resulted in the most severe growth
24      inhibition, regardless of the acidity of the SAR.
25           With such meager evidence, no clear conclusions can be drawn as to the ways in which the
26      effects of multiple airborne stressors could influence or be influenced by O3.
27
28      9.4.7  Interactions with Agricultural Chemicals
29           The review of interactions involving O3,  plants, and various agricultural chemicals
30      presented in the 1996 O3 AQCD (U.S. Environmental Protection Agency,  1996) remains a valid
31      assessment of our limited knowledge of these interrelationships. Our knowledge is largely based

        January 2005                             9-125        DRAFT-DO NOT QUOTE OR CITE

-------
 1      on the protection against O3 afforded to a range of crop species by applications of various
 2      chemicals, particularly fungicides, such as benomyl (benlate; methyl-l-[butylcarbamoyl]-2-
 3      benzimidazolecarbamate) and several carbamates and triazoles. A recent report has added
 4      azoxystrobin (AZO) and epoxyconazole (EPO) to the list (Wu and Tiedemann, 2002).  Foliar
 5      sprays of either AZO or EPO provided 50 to 60% protection against O3 injury to barley
 6      (Hordeum vulgare) leaves.  Both had similar modes of action involving stimulation of the levels
 7      of antioxidant enzymes such as SOD, ascorbate peroxidase, guaiacol peroxidase,  and catalase.
 8           In contrast, applications of herbicides have yielded variable results ranging  from increased
 9      sensitivity to protection from O3; the nature of the  effect is usually species- or cultivar-
10      dependent.  Although of less wide application, some plant growth retardants have also been
11      found to provide protection, but no insecticide appears to have been clearly shown to have
12      similar properties.
13           Despite the attraction of the use of permitted chemicals to provide crop protection, the
14      statement in the 1996 O3 AQCD (U.S. Environmental Protection Agency, 1996) is still valid:
15      "It is premature to recommend their use specifically for protecting crops from the adverse effects
16      of O3, rather than for their primary purpose."
17
18      9.4.8  Factors Associated with Global Climate Change
19           During the  last decade, interest in the effects of climatic change on vegetation has replaced
20      concerns  over the purported causes of forest decline and the effects of acidic deposition.  Two
21      specific components of climate change have been singled out as the foci of most of the research
22      activity:
23        •  the effects of increasing mean global CO2 concentrations in the lower atmosphere, and
24        •  the effects of increasing levels of surface-level irradiation by UV-B (the result of
            stratospheric O3 depletion).
25      In spite of the crucial role of temperature as a  climatic determinant (Monteith and Elston, 1993),
26      the effects of increasing mean global temperatures and their interactions with increasing CO2
27      levels in particular have received less attention.
28           All  of the biotic and chemical interactions with oxidants discussed in the preceding
29      sections may be modified by these climatic changes. However, research activities have largely

        January 2005                             9-126        DRAFT-DO NOT QUOTE OR CITE

-------
 1      focused on the two-way O3 x CO2 interaction.  Little if any experimental evidence exists related
 2      to three-way interactions such as O3 x CO2 x disease or O3 x CO2 x nutrient availability,
 3      although such interactions cannot be predicted from the component two-way interactions.
 4           Numerous reviews have appeared since the 1996 O3 AQCD (U.S. Environmental
 5      Protection Agency, 1996) dealing with the issues involved.  General reviews include
 6      publications of IPCC (1996); IPCC (2001); and UNEP (1993,  1999); the volume by Wellburn
 7      (1994); the volumes edited by Alscher and Wellburn (1994), De Kok and Stulen (1998), Singh
 8      (2000), and Yunus and Iqbal (1996); and papers by Idso and Idso (1994), Krupa and Groth
 9      (2000), Luo et al. (1999), Polle and Pell (1999), Poorter and Perez-Soba (2001), Runeckles
10      (2002), and Weubbles et al. (1999). Effects on agriculture and crop production, growth, and
11      metabolism have been reviewed by Groth and Krupa (2000), Rotter et al.  (1999), and Schnug
12      (1998); effects on forests have been reviewed by Bortier et al.  (2000); with focus on insect pests,
13      Docherty et al. (1997),  Karnosky et al. (2001b,a), McLaughlin and Percy  (1999), and Saxe et al.
14      (1998).
15           As background to the discussion of interactions with O3, it should be noted that the
16      increased levels of CO2 experienced since the mid-18th century are such that, without abatement
17      of the rates of increase, increased levels of from 540 to 970 ppm have been projected by the year
18      2100 (IPCC, 2001). Such increases in the concentration of CO2, the principal GHG released into
19      the atmosphere,  will inevitably lead to increased global mean temperatures, evidence for which
20      is already available from oceanic, icepack, and other records.  The latest estimates of the global
21      warming are for an increase in the range 1.4 to 5.8 °C over this century, in contrast to the 0.6 °C
22      rise experienced since 1900 (IPCC, 2001). However, considerable uncertainty  is associated with
23      such projections of future increases in global temperature.
24           The use  of elevated CO2 concentrations has been common practice for many years in the
25      production of many greenhouse crops. Much of our early knowledge of the effects of higher
26      than ambient CO2 levels on plant growth derives from this application, coupled with research of
27      plant physiologists on how CO2 concentrations affect the process  of photosynthesis. Information
28      available about effects  of increased CO2 levels on photosynthesis  and stomatal  function, in
29      particular, has provided the underlying bases for numerous process models that simulate plant
30      growth under  stress and in changed climates.
        January 2005                              9-127       DRAFT-DO NOT QUOTE OR CITE

-------
 1           Although simple O3 x temperature interactions were discussed in Section 9.4.4.2, the close
 2      linkage between global CO2 levels and global mean temperatures in the context of climate
 3      change requires that an assessment of the interactive effects with O3 should focus, as much as
 4      possible, on interactions involving all three factors.
 5
 6      9.4.8.1  Ozone-Carbon Dioxide-Temperature Interactions
 7           Idso and Idso (1994) reviewed several hundred reports published between 1982 and 1994
 8      on the effects of increased CO2 on plant growth and net photosynthesis. Their survey covered a
 9      wide range of temperate and tropical, herbaceous and perennial species, including coniferous
10      trees.  They concluded that, for responses to a 300-ppm increase in CO2, somewhat less than a
11      doubling of present-day levels, but somewhat greater than the 540 ppm lower limit suggested by
12      the IPCC (2001), averaged across all species:
13        •  light intensity had a negligible effect on net photosynthesis other than at limiting low
             intensities under which the CO2-driven enhancement was increased;
14        •  increased temperature tended to increase the CO2-driven enhancement of dry matter
             accumulation (growth) and net photosynthesis;
15        •  drought conditions tended to increase the CO2-driven enhancements of both growth and
             net photosynthesis, but increased salinity had  little effect;
16        •  mineral nutrient deficiency (especially of nitrogen) tended to increase the CO2-driven
             enhancement of growth; and
17        •  in the presence of air pollutants (especially SO2 and NOX), the CO2-driven enhancement
             of net photosynthesis tended to be increased.

18           It should be noted that the statement that CO2-enhanced growth increased with temperature
19      referred to total dry matter accumulation by the whole plant and not to the yield of grain, fruit, or
20      seed. Unfortunately, despite the existence of several reports at the time, the summary of
21      interactions with air pollutants contained only a single reference to O3, i.e., Pfirrmann and
22      Barnes (1993),  who reported surprisingly that a doubling of CO2 levels led to a 27% increase in
23      dry weight of radish (Raphanus sativus) but that the combination with  O3 led to a 77% increase.
24           The more recent reviews by Rudorff et al. (2000) and Olszyk et al. (2000) have addressed
25      CO2 x O3 interactions in detail, with the latter focusing on the implications for ecosystems. They
26      concluded that:

        January 2005                            9-128        DRAFT-DO NOT QUOTE OR CITE

-------
 1         •  the effects of both gases on stomatal closure were predominantly additive, with little
             evidence of interaction;
 2         •  increased photosynthesis resulting from elevated CO2 may be canceled by exposures to
             high O3 levels;
 3         •  foliar O3 injury is reduced by elevated CO2; and
 4         •  interactions between CO2 and O3 can affect storage carbohydrates, leaf free-radical
             metabolism, and carbon allocation to shoots and roots.

 5      Olszyk et al. also made specific note of the relative lack of information on below-ground effects.
 6           Much of the recently published information on the effects of increased CO2 and O3 levels is
 7      summarized in Table 9-12. Note that the table only lists the directions of O3-induced effects and
 8      any modifications of these effects resulting from elevated CO2, not their magnitudes. These
 9      directions are usually, but not necessarily, the same as the corollary effects of O3 on
10      CO2-induced responses.
11           The bulk of the available evidence clearly shows that, under the various experimental
12      conditions used (which almost exclusively employed abrupt or "step" increases in CO2
13      concentration, as discussed below), increased CO2 levels may protect plants from the adverse
14      effects of O3 on growth. This protection may be afforded in part by CO2 acting together with O3
15      in inducing stomatal closure, thereby reducing O3 uptake, and in part by CO2 reducing the
16      negative effects of O3 on Rubisco and  its activity in CO2-fixation. However, the situation with
17      regard to the combined effects of O3 and CO2 on stomatal functioning is not clear-cut. In spite of
18      the wealth of evidence supporting both CO2-induced and O3-induced short-term stomatal closure,
19      studies over long-term exposures, especially with tree species, have revealed little effect of
20      elevated CO2 on stomatal conductance. Reverse effects of O3 on stomata have also been noted,
21      as a result of O3-induced stomatal dysfunction after extended periods of exposure (Maier-
22      Maercker, 1998).
23           At the mechanistic level, Rubisco plays a key role in CO2-assimilation, and while both O3
24      and elevated CO2 per se can lead to reduced activity, CO2 can also reverse the O3-induced
25      inhibition of Rubisco activity and photosynthesis (Table 9-12).  However, in their review of the
26      possible mechanisms involved, Polle and Pell (1999) cautioned that Rubisco should not be
27      regarded "as a unique target for the interaction of the two gases." But it is clear from the bulk of
28

        January 2005                             9-129        DRAFT-DO NOT QUOTE OR CITE

-------
3
Table 9-12. Effects of Increased Carbon Dioxide on Ozone-Induced Responses of Plants at the Metabolic,
                               Physiological, and Whole-Plant Levels
to
o
o





VO
1
o


o
H
6
O
0
H
O
O
0
o
H
W
Co2 Effects:
Plant response O3 Response3
Biochemical/metabolic
Ascorbate peroxidase V
V
V
Catalase V
A
Chlorophyll V
V
Glutathione reductase V
O
O
Rubisco V
V
[A]
V



O3 Effects: V, Decrease; A, Increase; O, No Significant Effect.
A, Additive/Synergistic; T, Antagonistic/Ameliorative; O, No Significant Effect.
CO2 Modification"

V
OD
OD
OD
OD
T
V
T
OD
OD
T
V
V
T



Species

Wheat (Triticum aestivum)
Sugar maple (Acer saccharum)
Trembling aspen
(Populus tremuloides)
Wheat (T. aestivum)

Wheat (T. aestivum)
Potato (Solanum tuberosum)
Wheat (T. aestivum)
Sugar maple (A. saccharum)
Aspen (P. tremuloides)
Soybean (Glycine max)
Wheat (T. aestivum)
Trembling aspen
(P. tremuloides)
Sugar maple (A. saccharum)



Facility0

CSTR, P
CEQP
FACE, G
CEQP
CEQP
OTC, G
OTC, G
CSTR, P
CEQP
FACE, G
OTQP
OTC, G
FACE, G
CEQP



Reference

Rao etal. (1995)
Niewiadomska et al. (1999)
Wustman etal. (2001)
McKeeetal. (1997b)
Niewiadomska et al. (1999)
Ommenetal. (1999);
Donnelly et al. (2000)
Donnelly et al. (200 la)
Rao etal. (1995)
Niewiadomska et al. (1999)
Wustman etal. (2001)
Reid etal. (1998)
McKee et al. (2000)
Noormets et al. (2001)
Gaucher et al. (2003)




-------
3
Table 9-12 (cont'd). Effects of Increased Carbon Dioxide on Ozone-Induced Responses of Plants at the
                         Metabolic, Physiological, and Whole-Plant Levels
to
o
o





VO
1


o
H
6
O
0
H
O
O
0
o
H
W
O3 Effects: V, Decrease; A, Increase; O, No Significant Effect.
Co2 Effects: A, Additive/Synergistic; T, Antagonistic/Ameliorative; O, No Significant Effect.
Plant response O3 Response3 CO2 Modification6
Biochemical/metabolic (cont'd)
Rubisco activity V V
V T
V [T]
V T
Superoxide dismutase O OD
O OD
Physiological
Stomatal conductance V A
V T
V T
V [A]
A - OD T - OD
V A



Species

Soybean (G. max)
Wheat (T. aestivum)
Wheat (T. aestivum)
European beech
(Fagus sylvatica)
Wheat (T. aestivum)
Sugar maple (A. saccharum)

Radish (Raphanus sativus)
Soybean (G. max)
Bean (Phaseolus vulgaris)
White clover (Trifolium
repens) (O3-sensitive)
White clover (T. repens)
(O3-tolerant)
Tomato (Lycopersicon
esculentum)



Facility0

OTC,P
CEC,P
OTC, G
CEC,P
CEC,P
CEC,P

CEC,P
OTC, G
OTC,P
CSTR, P
CSTR, P
CEC,P



Reference

Reid etal. (1998)
McKeeetal. (1995)
McKee et al. (2000)
Liitz et al. (2000)
McKeeetal. (1997b)
Niewiadomska et al. (1999)

Barnes and Pfirrmann (1992)
Mulchi etal. (1992)
Heagle et al. (2002)
Heagle etal. (1993)
Heagle etal. (1993)
Hao et al. (2000)




-------
3
Table 9-12 (cont'd). Effects of Increased Carbon Dioxide on Ozone-Induced Responses of Plants at the Metabolic,
                                   Physiological, and Whole-Plant Levels
to
o
o










VO
1
to



o
H
1
O
o

o
H
O
o
H
w
o
^^
o
H
W
O3 Effects: V, Decrease; A, Increase; O, No Significant Effect.
Co2 Effects:
Plant response O3 response3
Physiological (cont'd)
Stomatal conductance V
(cont'd)
V

[V]
V
V

A

V
V
OD
OD
OD

OD


[A]

V





: A, Additive/Synergistic; T, Antagonistic/Ameliorative; O, No Significant Effect.
CO2 modification*

A

A

A
OD
O-A

OD

OD
OD
OD
OD
OD

OD


T

A





Species

Potato (S, tuberosum)

Wheat (T. aestivum)







Agropyron smithii
Koeleria cristata
Bouteloua curtipendula
Schizachyrium scoparium
Black cherry (Prunus serotina)

Green ash
(Fraxinus pennsylvanica)

Yellow poplar
(Liriodendron tulipiferd)
Trembling aspen
(P. tremuloides)




Facility0

OTC, G

CEQP

CEQP
OTC, G
CEQP

CEQP

CEQP
CEQP
CEQP
CEQP
CSTR, P

CSTR, P


CSTR, P

CEQP





Reference

Finnan et al. (2002)

Balaguer etal. (1995)
Barnes etal. (1995)
McKee etal. (1995)
Mulholland et al. (1997b)
Donnelly etal. (1998)

Tiedemann and Firsching (2000)

Volin etal. (1998)
Volin etal. (1998)
Volin etal. (1998)
Volin etal. (1998)
Loats and Rebbeck (1999)

Loats and Rebbeck (1999)


Loats and Rebbeck (1999)

Volin etal. (1998)






-------
3
Table 9-12 (cont'd). Effects of Increased Carbon Dioxide on Ozone-Induced Responses of Plants at the Metabolic,
                                   Physiological, and Whole-Plant Levels
to
o
o






VO
1
OJ


O
s
H
6
o
0
H
O
o
0
o
H
W
Co2 Effects:
Plant response O3 Response3
Physiological (cont'd)
Stomatal conductance O - V
(cont'd)
OD
[A]
Photosynthesis V
V
V
v
V
OD
OD
[V]
V

V



O3 Effects: V, Decrease; A, Increase; O, No Significant Effect.
A, Additive/Synergistic; T, Antagonistic/Ameliorative; O, No Significant Effect.
CO2 Codification"

V

OD
V
V
T
T
T
OD
OD
[T]
V
V

OD



Species



Red oak (Quercus rubra)
Durmast oak (Quercus petraed)
Radish (R. sativus)
Soybean (G. max)
Soybean (G. max)
Bean (P. vulgaris)
Tomato (L. esculentum)
Potato (S. tuberosum)

Wheat (T. aestivum)
Wheat (T. aestivum)





Facility0

FACE, G

CEQP
CEQP
CEQP
OTQP
OTQG
OTQP
CEQP
OTQG
OTQG
CEQP
OTQG
OTQP
CEQP
CEQP



Reference

Noormets et al. (2001)

Volin etal. (1998)
Broadmeadow et al. (1999)
Barnes and Pfirrmann (1992)
Booker etal. (1997)
Mulchi etal. (1992)
Heagle et al. (2002)
Hao et al. (2000)
Donnelly et al. (200 la)
Lawson etal. (200 Ib)
Barnes etal. (1995)
Donnelly et al. (2000)
Mulholland et al. (1997b)
Reid and Fiscus (1998)
Tiedemann and Firsching (2000)
Cardoso-Vilhena and Barnes (2001)




-------
3
Table 9-12 (cont'd). Effects of Increased Carbon Dioxide on Ozone-Induced Responses of Plants at the Metabolic,
                                   Physiological, and Whole-Plant Levels
to
o
o





VO
1



o
H
6
O
0
H
/->
Co2 Effects:
Plant response O3 Response3
Physiological (cont'd)
Photosynthesis (cont.) OD
V
o-v
[V]
V
V
OD
[V]
OD
V
V
OD
V
O3 Effects: V, Decrease; A, Increase; O, No Significant Effect.
A, Additive/Synergistic; T, Antagonistic/Ameliorative; O, No Significant Effect.
CO2 Codification"

OD
T
OD
T
T
T
OD
T
OD
T
T
OD
T
Species

Agropyron smithii
Koeleria cristata
Bouteloua curtipendula
Schizachyrium scoparium
Ponderosa pine
(Pinus ponderosa)
Scots pine (Pinus sylvestris)
Black cherry (P. serotina)
Green ash (F. pennsylvanica)
Yellow poplar (L. tulipifera)
Trembling aspen
(P. tremuloides)
European beech (F. sylvatica)
Red oak (Q. rubra)
Sugar maple (A. saccharum)
Facility0

CEQP
CEQP
CEQP
CEQP
CEQG
OTQ G
CSTR, P
CSTR, P
CSTR, P
CEQP
CEQP
CEQP
CEQP
Reference

Volinetal. (1998)
Volinetal. (1998)
Volinetal. (1998)
Volinetal. (1998)
Olszyketal. (2001)
Kellomaki and Wang (1997b);
Kellomaki and Wang (1997a)
Loats and Rebbeck (1999)
Loats and Rebbeck (1999)
Loats and Rebbeck (1999)
Volinetal. (1998)
Grams etal. (1999)
Volinetal. (1998)
Gaucher et al. (2003)
O
HH
H
W

-------
3
Table 9-12 (cont'd). Effects of Increased Carbon Dioxide on Ozone-Induced Responses of Plants at the Metabolic,
                                   Physiological, and Whole-Plant Levels
to
o
o





VO
1
0,


O
H
6
o
0
H
O
O
s
0
o
H
W
O3 Effects: V, Decrease; A, Increase; O, No Significant Effect.
Co2 Effects: A, Additive/Synergistic; T, Antagonistic/Ameliorative; O, No Significant Effect.
Plant response O3 Response3 CO2 Codification"
Physiological (cont'd)
Photorespiration V OD
V A
Growth, Yield
Total biomass V V
V T
V T
V T

[V] OD
V T
OD OD
V T

OD OD





Species

Soybean (G. max)
Wheat (T. aestivum)

Parsley (Petroselinum sativum)
Bean (P. vulgaris)

Soybean (G. max)

Alfalfa (Medicago sativa)
White clover (T. repens)
(O3-sensitive)
White clover (T. repens)
(O3-tolerant)
Tomato (L. esculentum)

Potato (S. tuberosum)





Facility0

OTQP
CEQP

CEQP
CEQP
OTQP
OTQG
CSTR, P
CEQP
CSTR, P
CSTR, P
CEQP
CSTR, P
OTQG





Reference

Booker etal. (1997)
McKeeetal. (1997b)

Cardoso-Vilhena et al. (1998)
Cardoso-Vilhena et al. (1998)
Heagle et al. (2002)
Mulchi etal. (1992)
Reinertetal. (1997)
Johnson etal. (1996a)
Heagle etal. (1993)
Heagle etal. (1993)
Hao et al. (2000)
ReinertandHo(1995)
Donnelly etal. (200 Ib);
Persson et al. (2003)





-------
3
Table 9-12 (cont'd). Effects of Increased Carbon Dioxide on Ozone-Induced Responses of Plants at the Metabolic,
                                   Physiological, and Whole-Plant Levels
to
o
o










VO
1
OJ
Oi



O
f?
H
1
O
o
0
H
O
o
0
o
H
W
O3 Effects: V, Decrease; A, Increase; O, No Significant Effect.
Co2 Effects:
Plant response O3 Response3
Growth, Yield (cont'd)
Total biomass (cont'd) V
V
V
v

V






[V]

V


V
[V]

V
V




A, Additive/Synergistic; T, Antagonistic/Ameliorative; O, No Significant Effect.
CO2 Codification"

0
T
T
T

T






V

OD


[V]
V

V
V




Species


Mustard (Sinapis alba)
Plantain (Plantago major)
Cotton (Gossypium hirsutum)

Wheat (T. aestivum)






Wheat (T. aestivum)

Wheat (T. aestivum)


Wheat (T. aestivum)
Timothy (Phleum pratense)

Agropyron smithii
Koeleria cristata




Facility0

OTC, G
CEC,P
CEC,P
OTC,P

CEC,P
OTC, G
OTC,P
CEC,P
OTC, G
CSTR, P
OTC, G
OTC, G

CEC,P


OTC, G
CEC,P

CEC,P
CEC,P




Reference

Lawsonetal. (200 la)
Cardoso-Vilhena et al. (1998)
Cardoso-Vilhena et al. (1998)
Booker, 2000)
Heagle etal. (1999)
Cardoso-Vilhena et al. (1998)
Fangmeier et al. (1996)
Heagle et al. (2000)
McKeeetal. (1997a)
Pleijel et al. (2000)
Rao etal. (1995)
Rudorffetal. (1996a)
Bender etal. (1999);
Mulholland et al. (1997a)
Cardoso-Vilhena et al. (1998)
Tiedemann and Firsching (2000)

Ewart and Pleijel (1999)
Johnson etal. (1996a)

Volin etal. (1998)
Volin etal. (1998)





-------
3
Table 9-12 (cont'd). Effects of Increased Carbon Dioxide on Ozone-Induced Responses of Plants at the Metabolic,
                                   Physiological, and Whole-Plant Levels
to
o
o






VO
1
3


o
^
H
6
O
0
H
O
O
s
0
o
H
W
Co2 Effects:
Plant response O3 Response3
Growth, Yield (cont'd)
Total Biomass (cont'd) V
V
O
V
V
[V]
A
O

[A]
V
V
V

0
V


O3 Effects: V, Decrease; A, Increase; O, No Significant Effect.
A, Additive/Synergistic; T, Antagonistic/Ameliorative; O, No Significant Effect.
CO2 Codification"

V
OD
OD
T
A
OD
V
OD

A
V
V
OD

OD
T


Species

Corn (Zea mays)
Bouteloua curtipendula
Schizachyrium scoparium
Ponderosa pine (P. ponderosa)
Birch (Betula pendula)
Black cherry P. serotina
Green ash (F. pennsylvanicd)
European ash
(Fraxinus excelsior)
Yellow poplar (L. tulipifera)
Sugar maple (A. saccharum)
Trembling aspen
(P. tremuloides)
(O3-tolerant clone}
(O3-sensitive clone)

Red oak (Q. rubra)
Durmast oak (Q. petraea)


Facility0

OTC, G
CEC,P
CEC,P
CEC, G
CEC,P
CSTR, P
CSTR, P
OTC, G

CSTR, P
CEC,P
CEC,P
OTC,P
OTC, G
OTC, G

CEC,P
CEC,P
OTC, G


Reference

Rudorffetal. (1996a)
Volinetal. (1998)
Volinetal. (1998)
Olszyketal. (2001)
Kytoviita et al. (1999)
Loats and Rebbeck (1999)
Loats and Rebbeck (1999)
Broadmeadow and Jackson (2000)

Loats and Rebbeck (1999)
Gaucher et al. (2003)
Volinetal. (1998)
Dicksonetal. (1998)
Dicksonetal. (2001)
Dicksonetal. (2001)

Volinetal. (1998)
Broadmeadow et al. (1999)
Broadmeadow and Jackson (2000)



-------
            Table 9-12 (cont'd). Effects of Increased Carbon Dioxide on Ozone-Induced Responses of Plants at the Metabolic,

                                               Physiological, and Whole-Plant Levels
to
o
o
oo
H

6
o


o
H

O
O3 Effects: V, Decrease; A, Increase; O, No Significant Effect.
Co2 Effects: A, Additive/Synergistic; T, Antagonistic/Ameliorative; O, No Significant Effect.
Plant response
Growth, Yield (cont'd)
Total Biomass (cont'd)

Seed/grain/fruit/tuber
yield







O3 Response3 CO2 Codification"

0 OD
0 OD
V T
V T
V T
V T
0 0
V T
[V] V
V [T]
Species Facility0

Aleppo pine (Pinus halepensis) CEC, P
Scots pine (P. sylvestris) OTC, G
Soybean (G. max) OTC, P
OTC, G
OTC, G
Bean (P. vulgaris) OTC, P
Tomato (L. esculentum) CSTR, P
Potato (S, tuberosum) OTC, G
OTC, G
Wheat (T. aestivum) OTC, G
OTC, G
CEC,P
OTC, G
OTC, G
OTC, G
Wheat (T. aestivum) (cont.) OTC, G
OTC, G
Reference

Kytoviita et al. (1999)
Broadmeadow and Jackson (2000)
Fiscusetal. (1997);
Mulchi et al. (1992);
Mulchietal. (1995)
Heagle et al. (2002)
ReinertandHo(1995)
Finnan et al. (2002)
Persson et al. (2003)
Bender etal. (1999);
Fangmeier et al. (1996);
McKeeetal. (1997a);
Mulchietal. (1995);
Mulholland etal. (1998a)
Rudorffetal. (1996b)
Fangmeier et al. (1996)
Mulholland et al. (1998b, 1998a)
o
HH
H
W

-------
3
Table 9-12 (cont'd). Effects of Increased Carbon Dioxide on Ozone-Induced Responses of Plants at the Metabolic,
                                   Physiological, and Whole-Plant Levels
to
o
o






VO
1
VO

o
s
H
6
O
0
H
O
O
0
o
H
W
Co2 Effects:
Plant response O3 Response3
Growth, Yield (cont'd)
Relative growth rate V
V
V
V
o
o
o
V
[V]
V
o
Specific leaf area-SLA V
V
V



O3 Effects: V, Decrease; A, Increase; O, No Significant Effect.
A, Additive/Synergistic; T, Antagonistic/Ameliorative; O, No Significant Effect.
CO2 Codification"

V
T
T
T
OD
OD
OD
V
V
V
OD
OD
T
OD



Species

Wheat (T. aestivum)

Agropyron smithii
Koeleria cristata
Bouteloua curtipendula
Schizachyrium scoparium
European ash (F. excelsior)
Trembling aspen
(P. tremuloides)
Red oak (Q. rubra)
Durmast oak (Q. petraea)
Scots pine (P. sylvestris)
Radish (R. sativus)
Soybean (G. max)




Facility0

CEQP
CEQP
CEQP
CEQP
CEQP
CEQP
OTQG
CEQP
CEQP
OTQG
OTQG
CEQP
OTQP
OTQG



Reference

Barnes etal. (1995)
Cardoso-Vilhena and Barnes (2001)
Volin etal. (1998)
Volin etal. (1998)
Volin etal. (1998)
Volin etal. (1998)
Broadmeadow and Jackson (2000)
Volin etal. (1998)
Volin etal. (1998)
Broadmeadow and Jackson (2000)
Broadmeadow and Jackson (2000)
Barnes and Pfirrmann (1992)
Reid etal. (1998)
Mulchi etal. (1992)




-------
            Table 9-12 (cont'd). Effects of Increased Carbon Dioxide on Ozone-Induced Responses of Plants at the Metabolic,

                                               Physiological, and Whole-Plant Levels
to
o
o
H

6
o


o
H

O

Plant response
Growth, Yield (cont'd)
Specific leaf area-SLA
(cont'd)







Root/shoot ratio




Co2 Effects:
O3 Response3

V
A
A
A
O
A
A
A
V
V
V
0
A
O3 Effects: V, Decrease; A, Increase; O, No Significant Effect.
A, Additive/Synergistic; T, Antagonistic/Ameliorative; O, No Significant Effect.
CO2 Codification"

A
A
A
A
O
O
A
A
V
A
T
OD
A
Species

White clover (T. repens)
(O3-sensitive)
White clover (T. repens)
(O3-tolerant)
Agropyron smithii
Koeleria cristata
Bouteloua curtipendula
Schizachyrium scoparium
Trembling aspen
(P. tremuloides)
Red oak (Q. rubra)
Radish (R. sativus)
Alfalfa (M. sativa)
White clover (T. repens)
(O3-sensitive)
White clover (T. repens)
(O3-tolerant)
Wheat (T. aestivum)
Facility0

CSTR, P
CSTR, P
CEC,P
CEQP
CEQP
CEQP
CEQP
CEQP
CEQP
CEQP
CSTR, P
CSTR, P
CEQP
Reference

Heagleetal. (1993)
Heagleetal. (1993)
Volinetal. (1998)
Volinetal. (1998)
Volinetal. (1998)
Volinetal. (1998)
Volinetal. (1998)
Volinetal. (1998)
Barnes and Pfirrmann (1992)
Johnson etal. (1996a)
Heagleetal. (1993)
Heagleetal. (1993)
McKeeetal. (1997a)
o
HH
H
W

-------
3
     Table 9-12 (cont'd). Effects of Increased Carbon Dioxide on Ozone-Induced Responses of Plants at the Metabolic,
                                             Physiological, and Whole-Plant Levels
to
o
o






VO
>^
^


o
H
6
O
0
H
O
Co2 Effects:
Plant response O3 Response3
Growth, Yield (cont'd)
Root/shoot ratio V
(cont'd)
O
O
0
0

Foliar injury A

A
A
A
A
A
Oo m < 015 nnm
O3 Effects: V, Decrease; A, Increase; O, No Significant Effect.
A, Additive/Synergistic; T, Antagonistic/Ameliorative; O, No Significant Effect.
CO2 Codification"

V

OD
OD
OD
OD

V

V
V
V
V
T

Species

Timothy (P. pratense)

Black cherry (P. serotina)
Green ash (F. pennsylvanica)
Yellow poplar (L. tulipifera)
Aspen (P. tremuloides)

Potato (S. tuberosum)

Bean (P. vulgaris)
Cotton (G. hirsutum)
Wheat (T. aestivum)
Trembling aspen
(P. tremuloides)
European beech (F. sylvatica)

Facility0

CEC,P

CSTR, P
CSTR, P
CSTR, P
OTC, G

OTC, G

OTC,P
OTC,P
CEC,P
OTC, G
FACE, G
CEC,P

Reference

Johnson etal. (1996a)

Loats and Rebbeck (1999)
Loats and Rebbeck (1999)
Loats and Rebbeck (1999)
Dicksonetal. (2001)

Donnelly etal. (200 Ib);
Persson et al. (2003)
Heagle et al. (2002)
Heagle etal. (1999)
Barnes etal. (1995)
Mulholland et al. (1997a)
Karnosky etal. (1999)
Wustman etal. (2001)
Grams etal. (1999)

O
H
W
b CO2-modifications of O3-effects resulting from ~2* present levels. (Trends are shown in brackets. Pronounced changes with ontogeny are, for example,
 indicated thus: OD-T.)
0 Exposure facilities used: CEC: controlled environment chambers; CSTR: continuously stirred tank reactors (Heck etal. 1978); FACE: free air CO2
 enrichment facilities; OTC:  open-top chambers. G: plants rooted in the ground; P: plants grown in pots.  All species are C3 except corn, Bouteloua
 and Schizachyrium.

-------
 1      the evidence in Table 9-12 that elevated CO2 levels can ameliorate the inhibition of growth
 2      caused by O3 in many species, although the precise balance among the mechanisms involved
 3      may well vary from species to species.  Three important caveats must be raised with regard to
 4      the findings presented in Table 9-12:
 5         •  the applicability of results from experiments with an abrupt (step) increase in CO2 level
              to understanding the consequences of the gradual increase in CO2 predicted for the
              troposphere over the next hundred years;
 g         •  the validity of the findings in several long-term studies (particularly with tree species)
              conducted using potted plants, because of possible added stresses imposed on their root
              systems relative to trees growing in the field; and
 7         •  the relevance to understanding the effects of climate change of studies focused solely on
              CO2 enrichment at current ambient conditions of temperature and precipitation patterns
              that provide no insights into possible interactive effects as these other climatic variables
              change concurrently with increasing CO2 (IPCC, 2001).

 8           The first caveat concerns the distinctly different natures of the exposures to O3 and CO2
 9      experienced by plants in the field. Changes in the ambient concentrations of these gases have
10      very different dynamics. In the context of climate change, CO2 levels increase relatively  slowly
11      and may change little over several seasons of growth.  On the other hand, O3 presents a
12      fluctuating stress with considerable hour-to-hour and day-to-day variability (Polle  and Pell,
13      1999). Almost all of the evidence presented in Table 9-12 comes from experimentation
14      involving plants grown from the outset in,  or subjected to, an abrupt or step increase to a higher
15      (more or less double), steady CO2 concentration. In contrast, the O3 exposure concentrations
16      usually varied from day to day. Luo and Reynolds (1999), Hui et al. (2002), and Luo (2001)
17      noted the difficulties in predicting the likely effects of a gradual CO2 increase from experiments
18      involving a step increase or those using a range of CO2 concentrations.  Indeed, although  using
19      the much accelerated time-scale of an 80-day growing season, Hui et al. (2002) clearly showed
20      significant differences between the rates and magnitudes of various physiological and growth
21      responses of plantain (Plantago lanceolatd) to CO2 between gradual and step increase
22      treatments.  The authors concluded that, even though there were major differences in most of the
23      parameters studied between the gradual and step treatments, "the convergence of the measured
24      parameters at the end of the experiment provides some encouragement for the applicability of
        January 2005                             9-142        DRAFT-DO NOT QUOTE OR CITE

-------
 1      step-type experiments in the field; however, the study suggests caution in interpreting early
 2      results from short-term studies."
 3           In long-term studies, the matter of photosynthetic acclimation to elevated CO2 levels has to
 4      be considered.  Lawlor and Keys (1993) define acclimation in terms of long-term (days, weeks),
 5      irreversible physiological changes, in contrast to regulation, which relates to more rapid
 6      (minutes, hours), reversible changes.  Each may be positive or negative, but many studies
 7      indicate that, while positive acclimation to elevated CO2 levels initially led to enhanced
 8      photosynthesis and growth, negative acclimation ultimately ensued and reduced CO2
 9      assimilation and growth rates. However, the consensus from recent studies and reviews is that
10      such negative acclimation is most likely to occur in situations in which plants are grown under
11      some additional stress, induced, e.g., by limitations to growth posed by lack of resources such as
12      water or nutrients. The meta-analysis by Curtis (1996) revealed that slow or little negative
13      acclimation was noted in studies on unstressed tree species with unhindered opportunities for
14      root growth and development, a view originally suggested by Arp and Drake (1991) and largely
15      supported in the review by Eamus (1996). A nonwoody perennial, the rhizomatous wetland
16      sedge, Scirpus olneyi, grown in its natural environment with no edaphic limitations showed no
17      negative acclimation after 4 years; in fact, photosynthetic capacity increased by 31% (Arp and
18      Drake, 1991). No negative acclimation of well-watered, field-grown Ponderosa pine (Pinus
19      ponderosd) trees was observed by Tissue et al. (1999) after 6 years of growth at 2x-ambient CO2
20      levels. Gifford and Morison (1993) have summarized the situation thus:  "Where the aerial or
21      root environment for a plant is restricted (as with inter-plant competition, for example), positive
22      feedback is limited and adjustments to the changed resource input balance under  high CO2 can
23      include 'down-regulation' of leaf photosynthesis rate as an integral part of a positive growth
24      response."
25           The influence of other environmental stress factors is borne out by several long-term tree
26      studies.  After 3 years in 565 ppm CO2 in the Duke Forest free-air CO2 enrichment (FACE)
27      facility in North Carolina, maturing loblolly pine (Pinus taedd) trees showed only a marginal
28      CO2-induced carbon  gain if grown on a nutritionally moderate site, but zero gain if grown on a
29      nutritionally poor site (Oren et al., 2001).  This is in sharp contrast to the substantially increased
30      initial growth rates in elevated CO2 reported by DeLucia et  al. (1999), but it is supported by the
31      observations  of Tognetti et al. (2000) on five Mediterranean tree species growing for many years

        January 2005                              9-143       DRAFT-DO NOT QUOTE OR CITE

-------
 1      adjacent to geothermal springs releasing CO2 to provide ambient levels averaging 700 ppm.
 2      No significant differences in radial growth of the oaks (Quercus cerris, Q. ilex, and
 3      Q. pubescens),  strawberry tree (Arbutus Unedo), and flowering ash (Fraxinus Ornus) could be
 4      detected between trees at the naturally enriched site and those at a nearby site exposed to normal
 5      ambient CO2 (-350 ppm). The authors concluded that limited availability of water and nutrients
 6      may have counteracted any positive effects of CO2 on growth at the enriched site or that the trees
 7      had acclimated to the higher CO2 levels.
 8           Because the ameliorative effects of CO2 on responses to O3 (Table 9-14) were reported
 9      mostly in short-term studies involving an abrupt increase in CO2 level, it is appropriate to ask
10      whether this amelioration is likely to persist to a time when the ambient CO2 concentration is
11      relatively stable at such levels. Regardless of any negative acclimation due to resource
12      limitations that may occur in the interim, steadily rising CO2 levels may well lead to natural
13      selection and genetic change. Nevertheless, it seems reasonable to expect that the amelioration
14      of O3 impact at elevated CO2 levels will be maintained in many situations, but the negative
15      acclimation that will probably occur in situations where other resources become limiting will
16      reduce the degree of protection.
17           Another caveat regarding the validity of some of the  observations in Table 9-12 is related
18      to the matter of stress-induced negative acclimation to elevated CO2 and concerns results
19      obtained using  potted plants. Although much of the recent information on CO2 effects has come
20      from experiments with plants rooted in the ground, more than half of the studies listed in
21      Table 9-12 used potted plants, whether in controlled environment and greenhouse chambers or in
22      OTCs.  The recent meta-analysis of data on the effects of elevated CO2 on soybean (Glycine
23      max) physiology and growth by Ainsworth et al. (2002) revealed a threefold smaller stimulation
24      of seed yield in pot-grown than in field-grown plants, even when large (> 9 L) pots were used.
25      Loats and Rebbeck (1999) noted that their inability to observe any CO2-induced increases in the
26      root/shoot ratios in seedlings of three broad-leaved tree species (Table 9-13) may have resulted
27      from their use of potted plants. The use of potted plants was a confounding factor in the studies
28      of Taylor et al.  (2001) of the differences in leaf growth of poplar (Populus) hybrids between
29      plants exposed  to elevated CO2 in controlled environment chambers (potted plants), OTCs or a
30      FACE facility.  Eamus (1996) has suggested that any long-term observations of reduced stomatal
31      conductance are almost universally a consequence  of pot-based trials at elevated CO2 with

        January 2005                             9-144       DRAFT-DO NOT QUOTE OR CITE

-------
 1      restricted root growth, in contrast to the lack of such a decline observed with trees rooted in the
 2      ground.  Although the majority of the cases cited in Table 9-12 indicate that O3 and CO2 act
 3      additively or synergistically in causing stomatal closure, there are numerous exceptions.
 4           Any reduction in stomatal aperture has consequences other than merely restricting O3
 5      uptake and the exchange of other gases.  In particular, the rate of transpiration is reduced and,
 6      while this tends to increase water-use efficiency, it may also lead to decreased mineral uptake,
 7      which could adversely impact growth over extended periods. Furthermore, less transpiration
 8      also means less evaporative cooling and an increase in leaf temperature, independent of any
 9      change in mean air temperatures.
10           Hence, the final caveat regarding Table 9-12 concerns the interactions of O3 and CO2 with
11      other climatic variables, especially mean temperature. In light of the key role played by
12      temperature in regulating physiological processes and modifying plant response to increased
13      CO2 levels (Long, 1991; Morison and Lawlor, 1999) and the knowledge that relatively modest
14      increases in temperature may lead to dramatic consequences in terms of plant development
15      (Lawlor,  1998), it is unfortunate that much of the large investment in time and resources spent
16      on recent studies  of the effects of climate change on vegetation have gone into investigations
17      limited to increasing our knowledge of the effects of higher levels of CO2 at current ambient
18      temperatures.
19           Some attention is now being paid to investigating the concurrent effects of CO2 increases
20      and warming (recently reviewed by Rowland-Bamford, 2000 and Morison and Lawlor, 1999),
21      but the observed interactive effects  on plant growth are inconsistent. For example, a FACE
22      study with ryegrass (Lolium perenne) ahowed that increased temperatures (provided by infrared
23      heaters) reduced the dry matter gain resulting from increased CO2 levels (Nijs et al.,  1996). The
24      field studies by Shaw et al. (2002) on a California annual grassland dominated by the grasses
25      Avena barbata and Bromus hordeaceus and the forbs Geranium dissectum and Erodium botrys
26      involved free-air  increased CO2 as well  as increased temperature, precipitation, and N supply.
27      Not only  did increased temperature reduce CO2-stimulated net primary productivity  (NPP), but
28      increased CO2 itself, combined with other factors, was found to be able to cause reduced NPP.
29           There have  been several investigations of effects on wheat (Triticum aestivum). Batts et al.
30      (1997) used plastic tunnels to create temperature gradients and maintain elevated CO2 levels
31      over field-grown  wheat and found that, in each  of 4 years of study, a temperature rise of-1.5 °C

        January 2005                              9-145       DRAFT-DO NOT QUOTE OR CITE

-------
 1      consistently canceled the growth and yield increases caused by a doubling of the CO2 level
 2      above ambient. Similar findings were reported by van Oijen et al. (1999) and van Oijen and
 3      Ewart (1999) in OTC field studies. Half of the chambers were cooled 1.6 to 2.4 °C below the
 4      uncooled chambers, to cancel out the normal temperature increase over ambient, due to the
 5      so-called "chamber effect" (usually a 1 to 3 °C increase above ambient temperature (Heagle
 6      et al., 1988).  Although temperature had no effect on CO2-enhanced assimilation rates, the
 7      CO2-enhanced growth and grain yields observed in the cooled chambers were effectively
 8      canceled out in the warmer chambers. The authors attributed this effect to accelerated
 9      phenology, a shorter period for grain filling, and a lower leaf area index (LAI; total leaf area per
10      unit ground area) in the warmer chambers. Wheeler et al. (1996) observed that the benefit to
11      wheat of doubling the CO2 level was  offset by a mean seasonal increase of only 1 to 1.8 °C.
12      With the continuing use of OTCs for field research, Runeckles (2002), has suggested that the
13      temperature rise due to the chamber effect in OTCs should be exploited (and measured) as a
14      means of exploring temperature x CO2 as well as temperature * CO2 * O3 interactions.
15          An indirect affirmation of the importance of temperature as a component of climate change
16      on wheat yield was provided by van Oijen and Ewart (1999), using two simulation models,
17      AFRCWHEAT2-O3 and LINTULCC (Ewart et al., 1999).  They analyzed data from the
18      ESP ACE-wheat program, which involved 25 OTCs experiments in 1994, 1995, and 1996 at nine
19      European locations (Jager et al.,  1999). Both models were able to predict control-treatment grain
20      yields closely (5.5 ±1.2 and 5.8 ±1.2 t.ha"1, respectively, versus the observed 5.9 ±1.9 t.ha"1),
21      and both indicated  a predominantly negative effect of temperature on the yield response to
22      increased CO2 (a 3  °C rise reduced the gain in yield from 30 to 14%). However, neither model
23      had an R2 > 0.3, indicating that the models included other sources of variability among the sites
24      than the climatic factors. The multiple linear regression developed by Bender et al. (1999) based
25      on the same data sets also included temperature as a highly significant covariant.  Both studies
26      are discussed more fully below.
27          Other studies, however, have found positive temperature-related growth effects, as
28      suggested by the early Idso and Idso (1994) analysis. In an OTC study using the perennial grass
29      Festucapratensis in which a temperature increase of 3 °C above ambient was combined with
30      CO2-enrichment to 700 ppm, both CO2 and temperature caused increases in total above-ground
31      biomass (Hakala and Mela, 1996).  Studies with  potato (Cao et al., 1994) and soybean (Ziska and

        January 2005                             9-146       DRAFT-DO NOT QUOTE OR  CITE

-------
 1     Bunce, 1997) using potted plants in controlled environment chambers also showed temperature-
 2     enhanced increases in growth in enriched CO2 atmospheres. Read and Morgan (1996) compared
 3     the effects of enriched CO2 and temperature on two grasses: cool-season Pascopyrum smithii
 4     and warm-season Bouteloua gracilis. In the latter (a C4 species), 750 ppm CO2 resulted in
 5     increased dry matter production at daytime temperatures as high as 35 °C, but in P. smithii
 6     (a C3 species), CO2-stimulated growth was greatest at 20 °C. However, the stimulation was
 7     progressively attenuated by increased temperature, so that at 35 °C, growth in 750 ppm was only
 8     one third of that in 350 ppm CO2 at 20 °C.
 9          Although the picture we have of temperature x CO2 interactions is inconsistent, Rowland-
10     Bamford (2000) has provided persuasive evidence that the nature of response to temperature in
11     the grain yield of crops with as different temperature optima as rice and wheat will depend upon
12     whether the change is above or below the temperature optimum.
13          But what if we add O3 as another variable? Unfortunately, there have been very few
14     studies of the three-way interaction. With the information available on CO2 x O3 interactions
15     (Table 9-12) and the limited information on temperature x O3 interactions (discussed in Section
16     9.4.4.2) simulation modeling can attempt to provide estimates  of O3 x CO2 x temperature effects,
17     but experimental observation is still required to validate the models.  The questions that need to
18     be answered are: if increased temperature can offset the gains in productivity afforded by
19     increased CO2 in important species such as wheat, and increased CO2 can offset the reductions in
20     productivity caused by O3, will increased temperature modify this protective effect.  And if so, in
21     what manner.
22          To date, the only information available appears to consist of the reports by van Oijen and
23     Ewart (1999) and Bender et al. (1999) referred to above. In the former's simulation studies, the
24     overall yield depression of wheat caused by O3 was found to be 7 ± 4% for both
25     AFRCWHEAT2-O3 and LINTULCC models versus an observed 9 ± 11%. The enhancements
26     due to CO2 were predicted to be 24 ± 9% and 42 ± 11%, respectively, which straddled the
27     observed 30 ± 22% gain. Based on the 13 experiments that included all four treatments (±O3,
28     ±CO2), an actual 10% yield loss due to O3 at ambient CO2 levels was reduced to a 4% loss by the
29     elevated CO2. The AFRCWHEAT2-O3 model predicted 7 and 4% losses, and LINTULCC
30     model predicted 8 and 5% losses due to the O3 and O3 + CO2 treatments. The actual and
31     simulated yield increases in response to CO2 increased further with increasing temperature, but

       January 2005                             9-147        DRAFT-DO NOT QUOTE OR CITE

-------
 1      although temperature had no discernible effect on the observed depression of yield caused by O3
 2      alone, both models suggested that the yield reduction was diminished both by higher
 3      temperatures and higher CO2 levels.
 4           The analysis of the ESPACE-wheat experiments by Bender et al. (1999) led to the
 5      following multiple linear regression:
 6
 7
 8      7=1004.6*** + 0.588***[C02]- 1.908**[O3]- 31.230***[r]+ 7.309[/]- 1448.423***[H2O], (9-5)
 9
10
11      where 7= grain yield, g-m~2; [CO2] = ppm CO2; [O3] = ppb O3, 12-h mean; [7] = °C;  [7] = light
12      intensity, MJ-nT2/day; and [H2O] is a dummy variable: well watered = 1; limited water
13      supply = 2. (***, p < 0.001; **, p < 0.01; the coefficient for/was not significant.)  With
14      R2 = 0.3983, adjusted for 258 degrees of freedom, a large part of the variability was still
15      unaccounted for by the five variables. However, this analysis suggests that CO2, O3,
16      temperature, and water-status are important codeterminants of wheat yield but assumes no
17      interactions.  Substitution in the model at summer light intensities and with well-watered plants
18      indicates that, at 20 °C, a doubling of CO2 levels to 700 ppm alone, would lead to a 29.5%
19      increase in yield, while 50 ppb O3 alone would decrease yield by 10.9%. With both gases at
20      those levels, the yield would only increase 20%, but with a concurrent temperature rise of 2 °C,
21      it would shrink to a 9.6% increase.
22           Both studies, therefore, indicate an amelioration of the effects of O3 by CO2, the magnitude
23      of which would be reduced at warmer temperatures. However, they relate to a single  crop whose
24      response to CO2 is temperature-sensitive. Information about other species in which the effects of
25      CO2 and temperature are additive are limited. However, Wolf and Van Oijen (2003) recently
26      described a model (LPOTCO) simulating the effects of changes in climatic variables,  CO2 and
27      O3 on tuber yield potential of irrigated potato (Solanum tuberosum cv. Bintje) over locations
28      within the European Union ranging from Finland to Italy. They noted that although increased
29      CO2,  O3, and light intensity were predominant controlling factors, increased temperature also
30      influenced potential yields substantially, with increases in northern latitudes (attributed to a
        January 2005                              9-148       DRAFT-DO NOT QUOTE OR CITE

-------
 1      longer growing season) but decreases in southern latitudes (attributed to decreased assimilate
 2      production).
 3           A clear understanding of the complex interactions of increased CO2 and temperature with
 4      O3 must await further experimentation or simulations. However, it seems likely that any CO2-
 5      induced amelioration of the adverse effects of O3 on aspects of growth other than seed or grain
 6      yield may be lessened or increased by increased temperature, depending upon the temperature,
 7      optima for the species, along the lines suggested by Rowland-Bamford (2000).
 8           Other crop simulation models which incorporate O3 and some of the various environmental
 9      factors, including elements of climate change, have been reviewed by Kickert et al. (1999) and
10      Rotter and van de Geijn (1999).  However, to date, the applications tend to have focused on
11      interactions of O3 with factors such as soil moisture or nutrient availability.
12           With forest trees, the situation has the added complexity of a perennial growth form and
13      the inevitability, over time, of subjection to additional environmental stresses such as nutrient-
14      limitation. Here, too, although numerous models of tree growth have been described, there
15      appear to have been few applications to interactions of O3 and factors of climate change.
16      Constable et al. (1996) used TREEGRO to model the growth of Ponderosa pine (Pinus
17     ponderosd) exposed to three  O3 levels (0.5x, l.Qx, and 2x ambient),  two levels of CO2 (ambient
18      and ambient + 200 ppm CO2), and two temperature regimes (ambient and ambient + 4 °C). Plant
19      growth was predicted to be decreased 1, 19, and 39% by the three levels of O3, respectively.
20      Increased CO2 reduced the loss at the highest O3 level to 7% but the  combination of elevated
21      CO2 with the higher temperature more than overcame the adverse effects of O3, leading to a 4%
22      increase, largely attributed to increased fine root mass.  The authors  suggested that, in relation to
23      the baseline conditions used in the simulations (Corvallis, OR), higher concentrations of CO2 and
24      O3 and a warmer climate will have little impact on total-tree growth, but they noted the
25      importance of undertaking multiple stress studies in order to be able  to make accurate forecasts
26      of the impact of such changes on forest trees.
27           More recently, Constable and Friend (2000) compared the capabilities  of six published
28      process-based models (CARBON, ECOPHYS, PGSM, TRE-BGC, TREEGRO, and W91) for
29      simulating tree response to elevated CO2, O3, and temperature. They concluded that although
30      these models were capable integrators of the effects of various environmental factors on
        January 2005                             9-149       DRAFT-DO NOT QUOTE OR CITE

-------
 1      individual processes such as photosynthesis, they were less reliable when extrapolating to
 2      growth.
 3           Although most of the research emphasis has been on simple CO2 x O3 interactions, a few
 4      isolated studies of interactions have involved O3, CO2, and biotic environmental factors. Heagle
 5      et al. (1994) observed that both O3 and CO2 tended to be additive in encouraging the growth of
 6      spider mite (Tetranychus urticae) populations on clover (Trifolium repens). Infection of wheat
 7      (Triticum aestivum) with leaf rust (Puccinia reconditd) sensitized the plants to O3 injury, but its
 8      severity was significantly reduced in elevated CO2 (Tiedemann and Firsching, 2000).  The
 9      effects of O3 and CO2 on mycorrhizal symbioses was studied by Kytoviita et al. (1999) who
10      found that CO2 did not ameliorate the adverse effects of O3 on the root growth of Aleppo pine
11      (Pinus halepensis) and birch (Betulapenduld). In another study with Aleppo pine, Kytoviita
12      et al. (2001) noted that both O3 and elevated CO2 reduced mycorrhiza-induced N-uptake by the
13      roots. In Scots pine (Pinus sylvestris\ Kasurinen et al. (1999) observed transient effects of
14      elevated CO2 and O3 on root symbiosis, but none of the effects persisted over the 3 years of
15      the study.
16           The soil water x O3 x CO2 interaction was experimentally investigated by Broadmeadow
17      and Jackson (2000) in Durmast oak (Quercuspetraea), European ash (Fraxinus excelsior), and
18      Scots pine (Pinus sylvestris).  No interactions were noted with ash and pine; but with oak,
19      elevated CO2 ameliorated and irrigation exacerbated the effects of O3, although the resultant
20      effects were essentially additive.
21           Booker (2000) noted that soil nitrogen levels interacted only slightly with O3 and CO2 in
22      determining the composition of cotton (Gossypium hirsutum) leaves and roots.  Carbon dioxide
23      reversed the inhibition of leaf growth caused by O3, but increased N-fertility tended to reduce
24      this reversal.
25           Because of the small number of studies of possibly significant interactions of three or more
26      environmental factors, it impossible to draw any sweeping conclusions as to how O3, in the
27      context of global climate change, may affect relationships among plants and insects, diseases,
28      and symbionts or among plants and nutrients or other air pollutants.  The only interaction that
29      has  some  degree of general support is the amelioration of adverse O3 effects by elevated CO2.
30
        January 2005                              9-150       DRAFT-DO NOT QUOTE OR CITE

-------
 1      9.4.8.2  Ozone-UV-B Interactions
 2           As noted in the 1996 O3 AQCD (U.S. Environmental Protection Agency, 1996), depletion
 3      of stratospheric O3 by halofluorocarbons has resulted in increased intensities of ultraviolet-B
 4      (UV-B) radiation (280 to 320 nm wavelengths) at the earth's surface.  The situation is discussed
 5      more fully in Chapter 10.
 6           While stratospheric O3 depletion may result in increased surface UV-B irradiation,
 7      absorption of UV-B is a property of the O3 molecule regardless of its location; and surface UV-B
 8      flux is, therefore, also reduced by O3 in the troposphere.  Although only about 10% of the total
 9      atmospheric O3 column occurs in the troposphere (Fishman et al., 1990), it contributes a
10      disproportionately greater absorption effect than  stratospheric O3 because the UV radiation
11      penetrating the troposphere becomes increasingly diffuse as it reaches the surface, with a
12      consequent increase in mean path length (Briihl and Crutzen, 1989).  Any benefits to vegetation
13      from reduced ambient O3 stress must, therefore, also be viewed in the context of possible
14      adverse effects due to increased UV-B irradiation.  There are, thus, two distinct types of possible
15      interactions between surface level O3 and UV-B radiation:
16         •  direct interactions involving simultaneous, sequential, or mixed exposures to O3 and
              UV-B stresses; and
17         •  effects on responses to UV-B itself resulting from changes in radiation intensity caused
              by changes in surface level O3 concentrations.
18      Only the first type of interaction is discussed here.  The second type of interaction has broad
19      implications for both health and welfare and focuses on the impacts of UV-V radiation per se.
20      It is dealt with  separately in Chapter 10, which includes a critical review of the experimental
21      difficulties faced in undertaking meaningful plant research in order to reach a clear
22      understanding of the effects of increased UV-B radiation, the evidence for both its adverse and
23      beneficial effects on plants, and the potential for  changes in ambient surface O3 levels to modify
24      these effects.
25           The most recent reviews specifically addressing the combined effects of tropospheric O3
26      and UV-B on plants are by Runeckles and Krupa (1994)  and Krupa and Jager (1996), although
27      the topic has also been included in several more general reviews of O3 effects and factors of
28      climate change such as those by Unsworth and Hogsett (1996), Krupa et al. (1998), Posthumus
29      (1998), Groth and Krupa (2000), and Krupa and Groth (2000).

        January 2005                              9-151       DRAFT-DO NOT QUOTE OR CITE

-------
 1           However, little new information has become available since Runeckles and Krupa (1994)
 2      noted that the scanty knowledge of the effects of UV-B and O3 combinations available at that
 3      time was derived solely from studies of soybean (Glycine max). Miller et al. (1994) observed no
 4      interaction and no effect of UV-B on yield, in contrast to a previous report by Teramura et al.
 5      (1990) using the same cultivar, Essex. More recently, in a study of the saltmarsh grass Elymus
 6      athericus subjected to reciprocal exposures to O3 and UV-B, van de Staaij et al. (1997) observed
 7      no interactive effects and no adverse effects of UV-B following 14-day exposures, even though
 8      an earlier report showed that longer  exposures to UV-B (65 days) could cause a 35% loss of
 9      biomass (Van De Staaij et al., 1993). However, in a study in which ambient, high altitude UV-B
10      levels were compared with near zero levels, at ambient or 2x-ambient levels of O3, interactions
11      involving the levels of antioxidants in Norway spruce (Picea abies) and Scots pine (Pinus
12      sylvestris) were reported by Baumbusch et al. (1998).  Schnitzler et al. (1999) subsequently
13      reported that O3-induced injury and  adverse effects on photosynthesis were more pronounced
14      with near zero UV-B levels, indicating an amelioration of the O3-response. A later study with
15      Scots pine (Zinser et al., 2000) revealed O3 * UV-B interactions at the gene expression and
16      biochemical levels. In contrast, Ormrod et al. (1995) reported that UV-B predisposed
17      Arabidopsis thalliana to injurious growth effects of O3.
18           At various organizational levels, Runeckles and Krupa (1994) identified several similarities
19      between plant response to O3 and UV-B, and at the level of gene expression there have recently
20      been several  reports of both similarities and distinctions.  Willekens et al.  (1994) reported similar
21      effects of O3, UV-B, and SO2 on the expression of antioxidant genes in Nicotiana
22      plumbaginifolia.  In parsley (Petroselinum crispum\ Eckey-Kaltenbach et al. (1994a) found that
23      O3 was a cross-inducer for both the UV-B-induced enhanced biosynthesis of flavonoids and the
24      pathogen-induced furanocoumarin phytoalexins, in keeping with the previously observed
25      O3-induction of fungal and viral defense reactions. In this regard, Yalpani et al. (1994) provided
26      evidence that, in tobacco (Nicotiana tabacum), O3 and UV-B acted similarly in increasing
27      disease-resistance via a salicylate-mediated enhancement of defense proteins. However,
28      subsequent work with tobacco led Thalmair et al. (1996) to conclude that exposure to UV-B did
29      not lead to the accumulation of pathogenesis-related proteins. In Scots pine (Pinus sylvestris),
30      although O3 is known to induce stilbene synthase and cinnamyl alcohol dehydrogenase, UV-B
        January 2005                              9-152       DRAFT-DO NOT QUOTE OR CITE

-------
 1     was found to enhance the former but suppress the latter, revealing an interaction at the level of
 2     gene expression (Zinser et al., 2000).
 3          In summary, the present base of information about possible interactions between increased
 4     UV-B  radiation and O3 is insufficient to draw any firm conclusions in terms of gross effects, but
 5     there is some evidence  of similarities in the effects of O3 and UV-B individually and of the
 6     mechanisms involved at the molecular level.
 7
 8     9.4.8.3 Interactions of Ozone with Multiple Climate Change Factors
 9          Despite of the need for experimental investigations of three-way or more complex
10     interactions among O3,  CO2, UV-B, temperature, and other climate change factors, few studies
11     have been reported, even without O3 as a factor.  In an isolated report, using tomato
12     (Lycopersicon esculentuni) seedlings, Hao et al. (2000) employed preexposure to UV-B (±CO2
13     enrichment) followed by exposure to O3 (± CO2 enrichment). They observed that CO2 more than
14     overcame the inhibition of photosynthesis caused by O3, but pretreatment with UV-B reduced
15     the resultant increase.
16          In view of the unexpected observations made in their grassland study of the combined
17     effects of CO2, temperature, precipitation,  and N-supply, Shaw et al. (2002) affirmed that:
18     "Ecosystem responses to realistic combinations of global changes are not necessarily simple
19     combinations of the individual factors." The addition of O3 to the list of variables results in
20     further complexity.
21          Although computer simulation modeling may ultimately lead to improved understanding of
22     these complex issues, to date, no such models appear to have been applied to these interactions,
23     possibly because of the scarcity of experimental data for parameterization.
24
25     9.4.9  Summary - Environmental Factors
26          Although O3 and other photochemical oxidants are phytotoxic, their actions on vegetation
27     may be modified by a host of biotic and abiotic factors in the environment; conversely, they may
28     modify plant response to these other factors. The extensive review of these biological, physical,
29     and chemical factors conducted for the 1996 O3 AQCD (U.S. Environmental Protection Agency,
30     1996)  concluded with a statement that our understanding was too fragmented to permit drawing
31     many general conclusions. With today's increased awareness of the need for more complete

       January 2005                             9-153        DRAFT-DO NOT QUOTE OR CITE

-------
 1      information on interactions, it is unfortunate that, in the interval since the 1996 criteria
 2      document, rigorous, systematic investigations of interactions have been rare, and most of the
 3      new information is as fragmented as before. This is inevitable partly in view of the vast scope of
 4      the possible interactions between O3 and other environmental variables.  But it is also an
 5      unavailable outcome of having the personal interests (and funding) of individual groups
 6      determine the studies to be undertaken, instead of a coordinated program of research focused on
 7      systematic investigations to improve our understanding our ability to assess the risks posed by
 8      photochemical oxidants to cultivated and natural vegetation.
 9           In the area of biotic interactions, new evidence with regard to insect pests and diseases has
10      done little to remove the uncertainties noted in the 1996 criteria document. Most of the large
11      number of such interactions that may affect crops, forest trees, and other natural vegetation have
12      yet to be studied. The trend suggested previously that O3 increases the likelihood and success of
13      insect attack has received some support from recent studies, but only with respect to chewing
14      insects. With the economically important group of sucking insects such as the aphids, no clear
15      trends have been revealed by the latest studies.  Hence, although it seems likely that some insect
16      problems could increase as a result of increased O3 levels, we are still far from being able to
17      predict the nature of any particular O3 plant insect interaction, its likelihood, or its severity.
18           The situation is a little clearer with respect to interactions involving facultative,
19      necrotrophic plant pathogens, with O3 generally leading to increased disease.  With obligate,
20      biotrophic fungal, bacterial, and nematode diseases, there are twice as many reports indicating
21      O3-induced inhibitions than enhancements. The frequent reports that infection by obligate
22      biotrophs reduces the severity of O3-induced foliar injury should not be interpreted as
23      "protection" because of the negative effects on the host plant of the disease per se. With obligate
24      biotrophs, the nature of any interaction with O3 is probably  dictated by the unique, highly
25      specific biochemical relationships between pathogen and host plant.  At this time, therefore,
26      although some diseases may become more widespread or severe as a result of exposure to O3, it
27      is still not possible to predict which diseases are likely to present the greatest risks to crops and
28      forests.
29           Several studies have indicated that  the functioning of tree root symbioses with mycorrhizae
30      may be adversely affected by O3, but there is also evidence that the presence of mycorrhizae may
31      overcome root diseases stimulated by O3 and that O3 may encourage the spread of mycorrhizae

        January 2005                              9-154        DRAFT-DO NOT QUOTE OR CITE

-------
 1      to the roots of uninfected trees. The latest studies, therefore, present no clearer picture of the
 2      likely nature of simple interactions of O3 and root symbionts, but in view of the importance of
 3      mycorrhizae as below-ground components of ecosystems, they are discussed more fully in
 4      Section 9.5.
 5           The few recent studies of the impact of O3 on intraspecific plant competition have again
 6      confirmed that grasses frequently show greater resilience than other types of plants.  In grass-
 7      legume pastures, the leguminous species suffer greater growth inhibition. And the suppression
 8      of Ponderosa pine seedling growth by blue wild-rye grass was markedly increased by O3.
 9      However, we are far from being able to predict the outcome of the impact of O3 on specific
10      competitive situations, such as successional plant communities or crop-weed interactions.
11           Light, a component of the plant's physical environment, is an essential "resource" whose
12      energy content drives photosynthesis and CO2 assimilation. It has been suggested that increased
13      light intensity may increase the sensitivity to O3 of light-tolerant species while decreasing that of
14      shade-tolerant species, but this appears to be an over-simplification with many exceptions.
15      Temperature affects the rates of all physiological processes based on enzyme-catalysis and
16      diffusion, and each process and overall growth (the integral of all processes) has a distinct
17      optimal temperature range. Although some recent field studies have indicated that O3 impact
18      significantly increases with increased ambient temperature, other studies have revealed little
19      effect of temperature.  But temperature is unquestionably an important variable affecting plant
20      response to O3 in the presence of the  elevated CO2 levels contributing to global climate change
21      (see below).  In contrast, evidence continues to accumulate to indicate that exposure to O3
22      sensitizes plants to low temperature stress by reducing below-ground carbohydrate reserves,
23      possibly leading to responses in perennial species ranging from rapid demise to impaired growth
24      in subsequent seasons.
25           Although the RH of the ambient air has generally been found to increase the adverse
26      effects of O3 by increasing stomatal conductance and thereby increasing O3 flux, abundant
27      evidence indicates that the ready availability of soil moisture  results in greater sensitivity to O3.
28      The partial "protection" against the adverse effects of O3 afforded by drought has been observed
29      in field experiments and modeled in computer simulations. There is also compelling evidence
30      that O3 can predispose plants to drought stress. Hence, the response will depend to some extent
31      upon the sequence in which the stresses occur, but, even though the nature of the responses is

        January 2005                             9-155        DRAFT-DO NOT QUOTE OR CITE

-------
 1      largely species-specific, successful applications of model simulations led to larger-scale
 2      predictions of the consequences of O3 x drought interactions.  However, it must be recognized
 3      that regardless of the interaction, the net result on growth in the short-term is negative, although
 4      in the case of tree species, other responses such as increased water use efficiency could be a
 5      benefit to long-term survival.
 6           Mineral nutrients in the soil, other gaseous air pollutants, and agricultural chemicals
 7      constitute chemical factors in the environment.  The evidence regarding interactions with
 8      specific nutrients is still contradictory. Some experimental evidence indicates that low general
 9      fertility increases sensitivity to O3, while simulation modeling of trees suggests that nutrient
10      deficiency and O3 act less than additively, but there are too many example of contrary trends to
11      permit  any sweeping conclusions. Somewhat analogously with temperature, it appears that any
12      shift away from the nutritional optimum may lead to greater sensitivity, but the shift would have
13      to be substantial before a significant effect on response to O3 was observed.
14           Interactions of O3 with other air pollutants have received relatively little recent attention.
15      The situation with SO2 remains inconsistent, but seems unlikely to pose any additional risk to
16      those related to the individual  pollutants. With the NO and NO2, the situation is complicated by
17      their nutritional value as a N source. In leguminous species, it appears that NO2 may reduce  the
18      impact of O3 on growth, with the reverse in other species, but the nature of the exposure pattern,
19      i.e., sequential or concurrent, also determines the outcome.  Much more investigation is needed
20      before we will be able to predict the outcomes of different O3-NO-NO2 scenarios. The latest
21      research into O3 x acid rain interactions has confirmed that,  at realistic acidities, significant
22      interactions are unlikely. A continuing lack of information precludes offering any
23      generalizations about interactive effects of O3 with NH3, HF, or heavy metals.  More evidence
24      has been reported that the application of fungicides affords some protective effects against O3.
25           Over the last decade,  considerable emphasis has been placed on research into O3
26      interactions with the components of global climate change:  increased atmospheric CO2,
27      increased mean global temperatures, and increased surface level UV-B radiation.  However,
28      most of these studies have tended to regard increased CO2 levels and increased mean
29      temperatures as unrelated phenomena. Experiments into the effects of doubled CO2 levels at
30      today's mean ambient temperatures are not paricularly helpful in trying to assess the impact of
31      climate change on responses to O3.  To date, the limited experimental evidence and evidence

        January 2005                              9-156        DRAFT-DO NOT QUOTE OR CITE

-------
 1      obtained by computer simulation suggest that in a 600+ ppm world, although the enriched CO2
 2      would more than offset the impact of O3 on responses as varied as wheat yield or the growth of
 3      young Ponderosa pine trees, the concurrent increase in temperature would reduce but probably
 4      not eliminate the net gain.  A similar decrease in the net gain resulting from the complete
 5      reversal by CO2 of the inhibition of photosynthesis caused by O3 has been reported for increased
 6      UV-B irradiation. However, these are preliminary results based on minimal data.
 7           In conclusion, although the increased use of computer simulations may be important in
 8      suggesting outcomes  of the many complex interactions of O3 and various combinations of
 9      environmental factors, the results obtained will only be as reliable as the input data used for their
10      parameterization. The data needed for good simulations can only come from organized,
11      systematic study.  For predicting the future, ignorance is as good as dependence on poor
12      simulations.
13
14
15      9.5   EFFECTS-BASED AIR QUALITY EXPOSURE-AND
16           DOSE-RESPONSE INDICES
17      9.5.1 Introduction
18           An index is needed that relates measured plant damage (i.e., growth) to ambient ozone
19      concentration over time. The index can provide both a consistency for reviewing exposure-
20      response effects in research, as well as a metric for developing a biologically-relevant air quality
21      standard that  protects ecological resources. The quantifying function over some time frame has
22      frequently been referred to as "dose-response" and "exposure-response".  The distinction being
23      where the pollutant concentration is measured:  "Dose" is the measure of the pollutant
24      concentration absorbed by the leaf over some time period, whereas "exposure" is the ambient air
25      concentration measured nearby the plant over some time period.
26           Plant ozone uptake from the ambient air (either rate or uptake or cumulative seasonal
27      uptake) is the ideal measure, because without ozone or its reactive product(s) reaching the target
28      tissue there is no effect. Uptake is controlled in part by stomata (see Section 9.3 for detailed
29      discussion). An uptake measure should integrate all those environmental factors that influence
30      stomatal conductance, e.g., temperature, humidity, soil water status. However, a direct measure
31      of the internal leaf concentration of ozone is technically difficult and thus uptake values are

        January 2005                             9-157       DRAFT-DO NOT QUOTE OR CITE

-------
 1      generally obtained with models that require species- and site-specific variables.  Because of this,
 2      a surrogate for uptake (i.e., exposure index) was sought early on using statistical summaries of
 3      ambient pollutant concentration over some integral of time (O'Gara, 1922; Lefohn and Benedict,
 4      1982; Lee et al., 1988; U.S. Environmental Protection Agency, 1986, 1992, 1996).
 5          An index of exposure that is biologically relevant must then consider those factors known
 6      to modify the plant response by altering ozone uptake (Hogsett et al., 1988; U.S. Environmental
 7      Protection Agency, 1996), including the temporal dynamics of exposure (e.g., concentration,
 8      frequency, duration), plant phenology (see Section 9.4), plant defense mechanisms (e.g.,
 9      antioxidants) (see Section 9.3), and site climate and soil factors (e.g., temperature, vpd, soil
10      moisture) (see Section 9.4).  In using these indices to develop air quality standards, the needs of
11      policy makers must also be considered, and those include simplicity and understandability
12      (Fairley and Blanchard, 1991). The development of such indices continues to be a challenge.
13
14      9.5.2   Summary of Conclusions from the Previous Criteria Document
15          The 1996 AQCD (U.S. Environmental Protection Agency, 1996a) focused primarily on
16      development of exposure indices, not flux, to quantify growth and yield effects in crops,
17      perennials and trees (primarily seedlings), and not foliar injury.  The testing of the adequacy of
18      these indices to order the measured responses of growth and/or yield in crops and tree species, as
19      seedlings, was accomplished through regression analyses of earlier exposure studies. No direct
20      experimental testing of the range of indices has yet been accomplished.  It was recognized that
21      these indices were only surrogates for O3 uptake or dose.  Their development focused on
22      consideration and inclusion of some, but not all, factors that affect O3 uptake and expression of
23      effects (e.g., Lee et al., 1988).  The 1996 document (U.S. Environmental Protection Agency,
24      1996a) drew a number of conclusions that built on even earlier conclusions (U.S. Environmental
25      Protection Agency, 1986, 1992). These conclusions are still valid today based on a review of
26      research published since  1996.
27          Studies prior to 1996, and after, indicate that the components of exposure, including
28      concentrations, temporal dynamics  (e.g., time of day of peak events), frequency of occurrence,
29      duration, and respite time, are integral to developing indices of exposure that relate to growth
30      response. Evidence from the few direct experimental studies of varying exposure components
31      indicate the importance of peak concentrations, occurrence, respite time and the importance of

        January 2005                             9-158       DRAFT-DO NOT QUOTE OR CITE

-------
 1      cumulative the concentrations (Musselman et al., 1983, 1986, 1994; Hogsett et al., 1885; U.S.
 2      Environmental Protection Agency, 1996a).
 3           Exposure duration influences the degree of plant response.  Single season, year-long or
 4      multi-year experimental results indicate that greater yield losses occurred when plants were
 5      exposed for the longer duration, and that a cumulative-type index was able to better describe the
 6      relationship between exposure and yield.  Those indices not considering duration, e.g., 7-h
 7      seasonal mean concentration index, single peak event index,  or the index that cumulates all
 8      concentrations (i.e., SUMOO), were unable to adequately describe the relationship. These single
 9      event or mean-type indices do not consider the role of duration of exposure and focus either only
10      on the peak event or put too much focus on the lower hourly  average concentrations (U.S.
11      Environmental Protection Agency, 1996a).
12           Higher hourly averaged concentrations had a greater affect on plant response. It was
13      concluded that cumulative indices that gave greater weight to higher concentrations relate well
14      with plant response (crops and tree seedlings) and order the treatment means in monotonically
15      decreasing fashion with increasing exposure, based on studies that apply two or more types of
16      exposure regimes with replicate studies of the same species.  These indices include, among
17      others: SUM06, W126, AOT40 (U.S. Environmental Protection Agency, 1996a).
18           No studies before or after 1996, have enabled a discrimination among the various
19      weighted, cumulative indices.  Various functional weighting  approaches were used, including
20      allometric, sigmoid or threshold weighting, and compared for best statistical fit of the plant
21      growth or yield data; but no one functional weighting was favored. For use as an air quality
22      standard, however, the need for simplicity, understandability, and ease of monitoring, favored
23      the cumulative threshold-weighted index (SUM06) (U.S. Environmental  Protection Agency,
24      1996a).
25           Since higher concentrations occur primarily during the  daylight hours, those indices that
26      differentially weight higher concentrations give greater weight to daylight hour concentrations.
27      This is important since stomatal conductance is usually greatest during the daylight hours,
28      compared to concentrations at night when conductance is usually minimal. Peak concentrations,
29      however, do not occur throughout the day; thus the timing of the peak concentration and
30      maximum  plant uptake is critical in determining plant response. An exposure index that
31      incorporated either the daily or seasonal temporal patterns of higher concentration occurrence

        January 2005                              9-159       DRAFT-DO NOT QUOTE OR CITE

-------
 1      with the temporal pattern of individual species' stomatal conductance was not reported in the
 2      prior 1996 review (U.S. Environmental Protection Agency, 1996).
 3           The relative importance of cumulative peak concentrations (> 0.10 ppm) versus cumulative
 4      mid-range concentrations (0.05-0.099 ppm) was questioned based on ambient field exposures of
 5      sensitive species. Although controlled experiments provided important evidence that the higher
 6      hourly average concentrations should be given greater weight than the mid-level values in
 7      developing indices, there was concern that under ambient conditions in the field the higher
 8      concentrations did not occur at the time of maximum plant uptake. This coincidence was
 9      considered to be the critical factor in determining peak concentration impacts on plants. Based
10      on the evidence at that time, it was not possible to conclude whether the  cumulative effects of
11      mid-range concentrations were of greater importance than those of peak  hourly average
12      concentrations in determining plant response (U.S. Environmental Protection Agency, 1996a).
13      No direct experimental studies had addressed this question prior to 1996, nor have any since.
14           The composite exposure-response functions for crops and tree seedlings were derived from
15      single and multi-year exposure studies that used modified or simulated ambient exposure
16      profiles. These profiles were typified by episodic occurrence of a large number of high O3
17      concentrations; and this type of pattern is prevalent in many but not all rural agricultural and
18      some forested areas in the United States.  Selecting a concentration value from these crop and
19      seedling response models may result in an over or underestimation of growth effects if applied to
20      regions of the country where a different type of exposure pattern is prevalent (U.S.
21      Environmental Protection Agency, 1996a). A multi-component index was suggested that
22      combined the peak-weighted, cumulative index with the number of occurrences of hourly
23      averaged concentrations > 0.10 ppm that might reduce the uncertainty associated with selecting
24      the exposure value for protection based on NCLAN type studies (Lefohn and Foley,  1992;
25      Musselman et al., 1994; U.S. Environmental Protection Agency, 1996a).  No direct experimental
26      studies addressed this question prior to 1996, nor have any since.
27           Since  1996, additional research has focused on the time of day when the higher hourly
28      average concentrations occur, the time of day of maximum plant uptake, and the diurnal
29      variability of plant defense mechanisms, and various suggestions  as to inclusion of these factors
30      in any one of the peak weighted cumulative exposure indices.  A much broader literature has
        January 2005                              9-160       DRAFT-DO NOT QUOTE OR CITE

-------
 1      focused since 1996 on developing indices using flux to relate plant response. These new
 2      developments are discussed in the sections that follow.
 3
 4      9.5.3 Use of Exposure Indices to Establish Exposure-Response Relationships
 5           Mathematical approaches for summarizing ambient air quality information in biologically
 6      meaningful forms that can serve as surrogates for dose for O3 vegetation effects assessment
 7      purposes have been explored for more than 80 years (O'Gara, 1922; U.S. Environmental
 8      Protection Agency, 1996). Several of the indices introduced have attempted to incorporate some
 9      of the biological, environmental, and exposure factors (directly or indirectly) that influence the
10      magnitude of the biological response and contribute to observed variability (Hogsett et al.,
11      1988). In the 1996 O3 AQCD (U.S. Environmental Protection Agency, 1996a), the exposure
12      indices were arranged into 5 categories, including (1) One Event, (2) Mean, (3) Cumulative,
13      (4) Concentration Weighted, and (5) Multi-component and were discussed in detail (Lee et al.,
14      1988). Figure 9-16 illustrates how several of the indices weight concentration and accumulate
15      exposure.
16           The indices were developed with knowledge of O3  modes of action, as well as the role of
17      the exposure components and their adequacy tested using earlier exposure studies, e.g., NCLAN
18      (U.S. Environmental Protection Agency, 1996a; Hogsett et al., 1988). Various components of
19      the exposure-response, including concentration, time of day, respite time, frequency of peak
20      occurrence, plant phenology, predisposition, etc., were weighted with various functions and
21      evaluated on their ability in ordering the regression of exposure vs. growth or yield response.
22      The statistical evaluations for each of these indices were accomplished using growth/yield
23      response data from many earlier exposure studies (e.g., NCLAN). This retrospective approach
24      was necessary because there have been very few studies  specifically designed to test the
25      goodness of fit  of the indices. The regression approach selects those indices that most properly
26      order and space the treatment means to optimize the fit of a linear or curvi-linear model. This
27      approach provides evidence for the best indices, albeit not as defensible as that from studies with
28      experimental designs and analyses that focus on specific components of exposure.
29           Most of the early retrospective studies reporting regression approaches use data from the
30      NCLAN program or  from Corvallis, OR (Lee et al., 1987; Lee et al.,  1988; Lefohn and Irving,
31      1988; Tingey et al., 1989) or use data collected in California (Musselman et al., 1988; U.S.

        January 2005                             9-161        DRAFT-DO NOT QUOTE OR CITE

-------
        Q.
        Q.
          0.15
          0.12-
          0.09-
o
Q.
X
LU

C
O
        o
        X
          0.06-
          0.03-
          0.00
                                                        8    10
n m

0 00-
PF>m 2ndHDM-»-
M-7 = 0.05 ppm


024681
Day
      Figure 9-16.  Diagrammatic representation of several exposure indices, illustrating how
                   they weight concentration and accumulate exposure,  (a) SUM06 - The
                   upper graphic illustrates an episodic exposure profile; the shaded area under
                   some of the peaks illustrates the concentrations greater than or equal to 0.06
                   ppm that are accumulated in the index.  The insert shows the concentration
                   weighting (0 or 1) function.  The lower portion of the  graphic illustrates how
                   concentration is accumulated over the exposure period, (b) SIGMOID - The
                   upper graphic illustrates an episodic exposure profile; the variable shaded
                   area under the peaks illustrates the concentration dependent weights that are
                   accumulated in the index. The insert shows the sigmoid concentration
                   weighting function. The mid-point of the sigmoid weighting scheme was
                   0.062 ppm. The lower portion of the graphic illustrates how concentration  is
                   accumulated over the exposure period, (c) 2ndHDM  and M-7 - The upper
                   graphic illustrates an episodic exposure profile.  The lower portion of the
                   graphic illustrates that the 2ndHDM considers only a single exposure peak
                   while the mean applies a constant exposure value over the exposure period.

      Source: Tinge etal. (1991).
1     Environmental Protection Agency, 1986).  These studies previously reviewed by EPA (U.S.

2     Environmental Protection Agency, 1992, 1996a) and were in general agreement and consistently

3     favored the use of cumulative peak-weighted exposure indices. Lee et al. (1987) suggested that

4     exposure indices that included all the data (24 h) performed better than those that used only 7 h

5     of data; this is consistent with the conclusions of Heagle et al. (1987) that plants receiving
      January 2005
                                      9-162
DRAFT-DO NOT QUOTE OR CITE

-------
 1      exposures for an additional 5-h/day showed 10% greater yield loss than those exposed for
 2      7-h/day. In a subsequent analysis using more of the NCLAN data, Lee et al. (1988) found the
 3      "best" exposure index was a phenologically weighted, cumulative index, with sigmoid weighting
 4      on concentration and a gamma weighting function as a surrogate for plant growth stage.  This
 5      index was the best statistical fit, but it depended upon more knowledge of species and site
 6      conditions that made specification of weighting functions difficult for general use.  This type of
 7      multi-function index did not meet the other criteria for developing an air quality standard, that
 8      being simplicity and understandability (Blanchard and Farley, 1999).
 9           The next best fits were the several indices which only cumulated and weighted higher
10      concentrations (e.g., SUM06, SUM08, AOT40, W126).  Amongst this group it was not possible
11      to distinguish a single best fit (Lee et al., 1988).  Similarly Lefohn et al., (1992) reported that it
12      was not possible to differentiate among the SUMOO, SUM06, SUM08, and W126 exposure
13      indices because the indices were highly correlated with one another in the experiment. Others
14      have reported similar results when attempting to identify optimum exposure indices (Musselman
15      etal., 1988).
16           Other factors, including predisposition time (Hogsett et al., 1988; McCool et al., 1988) and
17      crop development stage (Heagle et al., 1991; Tingey et al., 2002), contribute to variation in the
18      biological response and suggest the need for weighting O3 concentrations to account for
19      predisposition time and phenology. However, the roles of predisposition and phenology in
20      influencing plant response vary considerably with species and environmental conditions, so that
21      specification of a weighting function for general use in characterizing plant exposure is not
22      possible at this time.
23           In similar retrospective analyses, using data from the European Open-Top Chamber
24      Program, Finnan et al. (1997) confirmed that cumulative exposure indices which emphasize
25      higher O3 concentrations  are best related to plant response and that cumulative exposure indices
26      which use weighting approaches provide a better fit (U.S. Environmental Protection Agency,
27      1996a).
28           The main conclusions from 1996 regarding a biologically-relevant index based on ambient
29      exposure still hold true today.  No information has come forth in the interim that  alters those
30      conclusions, and in fact, some recent studies have further substantiated them. These key
31      conclusions can be restated as follows:

        January 2005                            9-163       DRAFT-DO NOT QUOTE OR CITE

-------
 I          •  ozone effects are cumulative;
 2          •  peak concentrations appear to be more important than lower concentrations in eliciting
               a response;
 3          •  plant sensitivity to O3 varies with time of day and crop development stage;
 4          •  exposure indices that accumulate the O3 exposure and preferentially weight the peaks
               yield better statistical fits to response than do the mean and peak indices;
 I          •  to avoid possible overestimation of vegetation effects when using NCLAN-type yield
               data in exposure-response models, the cumulative exposure index might be combined
               with number of hours of high hourly average concentrations over the time period.

 2          Following the 1996 review process (U.S. Environmental Protection Agency, 1996a,b), the
 3     EPA proposed an alternative form of the secondary NAAQS using a cumulative, concentration-
 4     weighted exposure index, the SUM06, to protect vegetation from damage (Federal Register,
 5     1997).  The EPA considered three specific concentration-weighted indices:  the threshold
 6     weighted SUM06, the AOT40, and the sigmoid weighted W126 exposure index (U.S.
 7     Environmental Protection Agency, 1996b). Both indices performed equally well in predicting
 8     the exposure-response relationships observed in the crop and tree seedlings studies conducted
 9     during the prior 20 years (Heck and Cowling,  1997). In the absence of research results that
10     differentiate the predictive power of these two forms, the EPA selected the SUM06 exposure
11     index as the form for the proposed secondary standard, recognizing its simplicity,
12     understandability, and ease of use (U.S. Environmental Protection Agency, 1996b).  A 3-month,
13     12-hour SUM06 exposure index of 26.4 ppm hr was proposed (U.S. Environmental Protection
14     Agency, 1996b). The value represented the concentration level that would protect 50% of the
15     crop species from a 10% yield loss. The composite response model included 49 studies
16     comprising 13 crop species and 5 locations across the U.S. where the crops are grown (U.S.
17     Environmental Protection Agency, 1996a).
18          European scientists took a similar approach as the United States in developing indices
19     describing growth and yield loss in crops and tree seedlings, using open-top chambers with
20     modified ambient exposures, but many fewer crop species were employed in the European
21     studies.  The European countries were seeking a scientific basis for control strategies to reduce
22     air pollution. They adopted the critical levels and loads approach, as per the UN Economic
23     Commission for Europe (UNECE). A critical level was defined as "the concentration of

       January 2005                             9-164       DRAFT-DO NOT QUOTE OR CITE

-------
 1      pollutant in the atmosphere above which direct adverse effects on receptors, such as plants,
 2      ecosystems, or materials may occur according to present knowledge" (United Nations Economic
 3      Commission for Europe, 1988).  Critical levels are set to prevent long-term injury and damage to
 4      the most sensitive elements of any ecosystem.  They are used to map and identify areas in
 5      Europe in which the levels are exceeded and that information is then used to plan optimized and
 6      effect-based abatement strategies. As used by the UNECE, they are not air quality standards in
 7      the US sense, but they have been used as targets for planning reductions in sulfur and nitrogen
 8      emissions to protect ecological resources.  The nature of the significant harmful effects is not
 9      specified in the original definition, which provides for different levels for different types of
10      harmful effect (e.g., visible injury or loss of crop yield). There are also different levels for crops,
11      forests, and semi-natural vegetation.  The caveat, "according to present knowledge," is important
12      because critical levels are not rigid; they are revised periodically as new scientific information
13      becomes available.  For example, the original critical level for O3 specified concentrations for
14      three averaging times, but further research and debate led to the current critical levels being
15      stated as the cumulative exposure (concentration x hours) over a cut-off concentration of 40 ppb
16      (AOT40) (Fuhrer et al., 1997). The "Level I" critical level was used in the 1990s to map areas of
17      exceedance, but analyses of many exposure studies led to the conclusion that the simple,
18      exposure-based level leads to over-estimation of the effects in some regions and
19      under-estimation in others (Fuhrer et al., 1997; Karenlampi and Skarby, 1996). The main
20      problem was that other environmental factors (vapor pressure deficit, water stress, temperature,
21      and light) and variation in canopy height altered O3 uptake and its effects.
22           A decision was made to work towards a flux-based approach for the critical level, with the
23      goal of modeling O3 flux-effect relationships for three vegetation types:  crops, forests, and
24      semi-natural vegetation.  Progress has been made in modeling flux (e.g., Grunhage and Jager,
25      2003; Ashmore et al., 2004a; Ashmore et al., 2004b) and the Mapping Manual is being revised
26      (Karlsson et al., 2003; Ashmore et al., 2004a; Ashmore et al., 2004b; Grennfelt, 2004).  The
27      revisions may include a flux-based approach for 3 crops: wheat, potatoes, and cotton; but,
28      because of lack of flux data, a cumulative, concentration-based (AOTx) exposure index will
29      remain for most crops, and for forests and semi-natural herbaceous vegetation (Ashmore et al.,
30      2004; Jaeger and Grunhage, 2004).
31

        January 2005                              9-165       DRAFT-DO NOT QUOTE OR CITE

-------
 1      9.5.4  Identifying Exposure Components That Relate to Vegetation Effects
 2           The efficacy of exposure indices to predict biological response requires that researchers
 3      identify a relationship between exposure components and effects, as well as those environmental
 4      and site factors that control pollutant uptake by the plant. These relationships were identified
 5      and discussed in the 1996 review (U.S. Environmental Protection Agency, 1996a).
 6      A significant, but in some instances, unquantifed role was identified for:  (1) concentration;
 7      (2) duration of exposure; (3) the diurnal and seasonal patterns of exposure, e.g., time of day of
 8      peak event, season of higher exposures, seasons of high precipitation and humidity, the
 9      frequency of occurrence of peak events and respite time  (peak to valley ratios); (4) plant
10      phenology; (5) plant canopy structure; (6) meteorological and site factors, e.g., light, humidity;
11      and (7) plant defense mechanisms.
12
13      9.5.4.1 Role of Concentration
14           A significant role of higher concentrations was established earlier, based on several
15      experimental studies (U.S. Environmental Protections Agency, 1996a). Recently Nussbaum
16      et al., (1995) and Yun and Laurence (1999b) have added support for the important role that peak
17      concentrations, as well as the pattern of occurrence, plays in plant response to O3. Based on air
18      quality data from 10 U.S. cities, three treatments of 4-week exposure to the same SUM06 value
19      were constructed by Yun and Laurence  (1999b).  They used the regimes to explore effects of
20      treatments with variable peak occurrence versus uniform peak occurrence during the exposure
21      period. The authors reported that the peak exposures were important and that the same SUM06
22      resulted in very different foliar injury.  Oksanen and Holopainen (2001) found that the peak
23      concentrations and the shape of the O3 exposure (i.e., duration of the event) were important in
24      foliar injury of white birch saplings, but growth reductions were found to be  more related to
25      cumulative exposure. Nussbaum et al. (1995) also found peak concentrations and the pattern of
26      occurrence to be critical in the measured response. The authors recommended that to describe
27      the effect on total forage yield, peak concentrations > 0.11 ppm must be emphasized by using an
28      AOT with higher threshold concentrations.
29           A greater role for higher concentrations affecting plant growth may be  inferred based on
30      recent air quality analyses for the Southern California area (Lee et al., 2003;  Tingey et al., 2004).
31      In the late 1960s and 1970s, extremely high O3 concentrations had impacted  the San Bernardino

        January 2005                              9-166       DRAFT-DO NOT QUOTE OR CITE

-------
 1     National Forest. However, over the past 15 plus years, significant reductions in the O3 exposure
 2     have occurred in the San Bernardino National Forest (Lee et al., 2003; Lefohn and Shadwick,
 3     2000; Davidson, 1993; Lloyd et al., 1989).  An illustration of this improvement in air quality is
 4     shown by the 37 year history of ozone air quality at a site in the San Bernardino Mountains
 5     (Figure 9-17) (Lee et al., 2003). The O3 exposure increased from 1963 to 1979 concurrent with
 6     increased population and vehicular miles, followed by a decline to present mirroring decreases in
 7     precursor emissions.  The pattern in exposure was evident in various exposure indices including
 8     the cumulative concentration weighted (SUM06), as well as maximum peak event (1-h peak),
 9     and the number of days having hourly averaged O3 concentrations > 95 ppb (i.e., the California
10     ozone standard). The number of days having hourly averaged O3 concentrations > 95 ppb
11     declined significantly from 163 days in 1978 to 103 days in 1997.  The changes in ambient
12     ozone air quality for the site were reflected in the changes in the frequency and magnitude of the
13     peak hourly concentration and the duration of the exposure (Figure 9-17). Considering the role
14     of exposure patterns in determining response, the seasonal and diurnal patterns in hourly O3
15     concentration did not vary appreciably from year to year over the 37-year period (Lee et al.,
16     2003).
17           The inference for a role of higher concentrations comes from both results of ground
18     measures of tree conditions on established plots and from model simulations. Across a broad
19     area of the San Bernardino National Forest, the Forest Pest Management (FPM) method of injury
20     assessment indicated an improvement of crown condition from 1974 to 1988; and the area of
21     improvement in injury assessment is coincident with an improvement of O3 air quality (Miller
22     and Rechel, 1999). A more recent analysis of forest changes in the San Bernardino National
23     Forest using an expanded network of monitoring  sites has verified significant changes in growth,
24     mortality rates, basal area, and species composition throughout the area since 1974 (Arbaugh
25     et al., 2003). A model simulation of ponderosa pine growth over the 40 year period in the
26     San Bernardino showed a significant impact of ozone exposure on tree growth and indicates
27     improved growth with improving air quality.  The improvement in growth was assigned to
28     improved air quality, but no distinction was made regarding the relative role of mid-range and
29     higher hourly concentrations, only that improved growth tracked both decreasing SUM06,
30     maximum peak concentration and number of days of hourly O3 > 95 ppb (Tingey et al., 2004).
31

       January 2005                             9-167        DRAFT-DO NOT QUOTE OR CITE

-------
•if

E I
LLJ &
        Q.
        E 0
       ,0) 2_x
       §*=
       ^ E
       D Q-
       w S
        TO _
        o> S~
       CL Q.
        ro Q-
       Q if)
             o
             in
             
-------
 1      9.5.4.2 Role of Duration
 2           Recent studies have called into question the period of time over which concentrations are
 3      accumulated and the form of the exposure index. Heagle and Stefanski (2000) reported that the
 4      form of the exposure index was important only for 24-h indices for which SUMOO provided the
 5      poorest fit. The authors reported that the SUMOO, SUM06, W95 (Lefohn and Runeckles, 1987),
 6      W126, and AOT40 produced similarly good fits of the foliage biomass data for 6-, 5-, and 4-h
 7      midday accumulating periods.  The study "pooled" data from San Bernardino (CA) and
 8      Riverside (CA) with data from Amherst (MA), Corvallis (OR), Kennedy Space Center (FL),
 9      Raleigh (NC), and Blacksburg (VA).  Ozone exposures were much higher at the two California
10      sites (indicated by high W126, SUM06, W95, and AOT40  values) in comparison to the other
11      locations. Because of the pooling of the data, the large number of high hourly average O3
12      concentrations that occurred at the California sites may have resulted in the exposure indices
13      being highly correlated with one another and made it difficult to identify one optimal index.
14           In another study in California, Arbaugh et al. (1998) reported that the SUMOO exposure
15      index performed better for describing visible injury than the SUM06, W126, number of hours
16      greater than or equal to 0.08 ppm, and the number of days between measurement periods.  These
17      exposure indices were originally developed and tested using only growth/yield data, not foliar
18      injury (U.S. Environmental  Protection Agency,  1996a).  This distinction is critical in comparing
19      the efficacy of one index to another.
20           The SUMOO exposure index is a surrogate for a long-term average concentration and
21      earlier growth response studies indicated that a long-term average was not adequate in predicting
22      effects primarily because it gives equal weight to all concentrations (U.S. Environmental
23      Protection Agency, 1996). However, for many locations in California, a large number of higher
24      hourly average concentrations occur and the SUMOO could be highly correlated with the
25      frequency of elevated hourly average concentrations and thus could be a good predictor of
26      vegetation effects. In Section 9.3, it was noted that research results are mixed on what causes an
27      effect in plants:  an event or the cumulation of events.
28
29      9.5.4.3 Patterns of Exposure
30           A significant factor in developing exposure indices is the temporal patterns of ozone
31      occurrence over a day, a month, a year, and seasonally overlaying the daily and seasonal

        January 2005                            9-169        DRAFT-DO NOT QUOTE OR CITE

-------
 1     temporal patterns of those influential climatic and site factors.  The coincidence of peak ozone
 2     and maximal stomatal conductance and detoxification processes are key to effecting plant
 3     growth response (Musselman and Minnick, 2000).
 4
 5     Daily Patterns
 6           Recent experimental studies with tree species have demonstrated the de-coupling of
 7     ambient O3 exposure, peak occurrence, and gas exchange, particularly in areas of drought
 8     (Panek, 2004).  The coincidence of maximal conductance and occurrence of higher
 9     concentrations are relevant to the time of day that exposure is cumulated in those indices that
10     cumulate and weight concentration.
11           The hours during the day over which ambient exposure is cumulated or effective dose is
12     estimated is a surrogate for the coincidence of these patterns of conductance and O3 occurrence.
13     A 12-hr daylight period for cumulating exposure was proposed following the 1996 review (U.S.
14     Environmental Protection Agency, 1996). An extensive review of the literature revealed that a
15     large number of species had varying degrees of nocturnal stomatal conductance (Musselman
16     and Minnick, 2000). Grulke et al. (2003) showed that the stomatal conductance at night for
17     ponderosa pine in the San Bernardino (CA) National Forest ranged from one tenth to one fourth
18     that of maximum daytime gas exchange.  In June, at the high-elevation site, 11% of the total
19     daily O3 uptake of pole-sized trees occurred at night. In late summer, however, O3 uptake at
20     night was negligible. Birch seedlings exposed to O3 at night show greater reductions in growth
21     than those exposed to O3 in daylight (Matyssek et al., 1995). Brassica rapa plants exposed to O3
22     during the day or night, show little significant difference in the amounts of injury or reduced
23     growth response to treatment; the conductance was 70 to 80%  lower at night (Winner et al.,
24     1989). Tissue biomass of ponderosa pine seedlings was significantly reduced when seedlings
25     were exposed to either daytime or nighttime episodic profiles (Lee and Hogsett 1999).  However,
26     the biomass reductions were much greater with daytime peak concentrations than with nighttime
27     peak concentrations.
28           Although conductance is lower at night than during the day for most plants, nocturnal
29     conductance can result in some measurable O3 flux into the plants and should be considered.
30     Nocturnal O3 flux also depends on the level of turbulence that intermittently occurs at night.
31     In addition, plants can be more susceptible to O3 exposure at night than during the daytime,

       January 2005                            9-170        DRAFT-DO NOT QUOTE OR CITE

-------
 1      because of lower plant defenses at night (Musselman and Minnick, 2000). Massman (2004)
 2      suggested that nocturnal stomatal O3 uptake accounted for about 15% of the cumulative daily
 3      effective O3 dose that was related to predicted injury.  Based on a review of the literature relating
 4      to nocturnal stomatal conductance, Musselman and Minnick (2000) recommended that any O3
 5      exposure index used to relate air quality to plant response should use the 24-h cumulative
 6      exposure period for both exposure-response and effective flux models.
 7           In general, stomatal conductance needs to be taken into account in developing indices
 8      (Panek et al., 2002). Stomatal conductances are linked to both diurnal and seasonal
 9      meteorological activity and soil/site conditions (e.g.,  soil moisture). Daily patterns of
10      conductance are often highest in mid-morning, whereas higher ambient O3 concentrations
11      generally occur in mid to late afternoon, when stomata are often partially closed and
12      conductances are lower. Total flux depends on atmospheric resistance and boundary layer
13      resistances, both of which exhibit variability throughout the day.  Several recent studies have
14      suggested that ponderosa pine trees in the southern and northern Sierra Nevada Mountains may
15      not be as sensitive to high O3 concentrations as to lower concentrations,  due to reduced O3
16      uptake during the period when the highest concentrations occur (Panek et al., 2002; Panek and
17      Goldstein 2001; Bauer et al., 2000; Arbaugh et al.,1998).  Panek et al. (2002) compared direct
18      measurements of ozone flux into a canopy of ponderosa pine and demonstrated a lack of
19      correlation of daily  patterns of conductance  and ozone occurrence, especially in the late season
20      drought period; and they concluded that a consideration of climate or season was essential,
21      especially considering the role of soil moisture and conductance/uptake.  In contrast, Grulke
22      et al. (2002) reported high conductance when O3 concentrations were high in the same species,
23      but different growing site conditions.  The uncoupling of conductance and higher ambient O3
24      concentration would hold true for more mesic environments as well as xeric landscapes.  The
25      longer term biological responses reported by Miller and Rechel (1999) for ponderosa pine in the
26      same region and the general reduction in recent years in ambient O3 concentrations, suggest that
27      conductance alone may not be a sufficient indicator of potential vegetation injury or damage.
28           The generalized models of stomatal conductance may  provide a means to link patterns of
29      O3 occurrence with  climatic and site factors  that affect O3 uptake to some degree, provided
30      conductance is modeled by regions of similar seasonal moisture and by similar canopy structure
31      (e.g., Emberson et al., 2000a,b; Grunhage et al., 2000; Massman, 2004)

        January 2005                              9-171       DRAFT-DO NOT  QUOTE OR CITE

-------
 1      Seasonal Patterns
 2           Several of the recent studies measuring O3 flux to pine canopies also reported on the
 3      importance of seasonal patterns in relating exposure to response (Panek, 2004). These seasonal
 4      patterns can be early vs late season occurrence of higher O3 concentrations, reflecting climate
 5      and precursor availability. The patterns also reflect seasonal drought and the role of soil
 6      moisture plays in stomatal conductance and O3 uptake. Recently, studies have looked directly at
 7      this linkage.  Panek et al. (2002) compared direct measurements of ozone flux into a canopy of
 8      ponderosa pine with a number of exposure indices, demonstrated a lack of correlation especially
 9      in the late season drought period, and concluded that a consideration of climate was essential,
10      especially soil moisture.  They suggested that a better  metric for a  seasonally drought-stressed
11      forests would be one that incorporates forest physiological activity, either through mechanistic
12      modeling, by weighting ambient O3 concentrations by stomatal conductance, or by weighting O3
13      concentrations by site moisture conditions. Panek (2004) demonstrated a decoupling of O3
14      exposure and uptake seasonally as well, via seasonal drought influence. Maximum O3 uptake
15      occurred at the beginning of the season and in the winter, whereas  the pines were nearly dormant
16      during August-September.
17           Using TREGRO, a  process-based model, Tingey et al. (2004) simulated long-term growth
18      of ponderosa pine over a  37-year period. The simulation showed high correlation between O3
19      exposure and O3-induced reductions in tree growth (R2 = 0.56).  The scatter about the line
20      however indicates that other factors besides O3 are required to describe the  association between
21      exposure and response. Incorporating annual temperature and precipitation increased the R2 to
22      0.67. In keeping  with the observations of Panek (2004) on the decoupling of peak O3 occurrence
23      and maximal conductance, the remaining unexplained variation is  attributed to differences in
24      timing of peak O3 uptake and  peak O3 exposure over the years.
25
26      9.5.4.4  Frequency of Occurrence of Peak Concentrations
27           Several earlier studies demonstrated the greater effect of episodic occurrence of O3 peaks
28      compared to daily peak events (U.S. Environmental Protection  Agency, 1996a); and, since the
29      last review, a few studies have corroborated the importance of this pattern in growth response.
30      Kollner and Krause (2003) reported that, under equal exposure conditions, the most pronounced
31      effects on the yield  of sugar beet and soybeans occurred with those regimes that emphasized the

        January 2005                              9-172       DRAFT-DO NOT QUOTE OR CITE

-------
 1      episodic occurrence of peak events.  Nussbaum et al. (1995) compared the effects of different
 2      patterns of peak occurrences with similar AOT40 values and reported a stronger effect on total
 3      forage yield from the episodic treatment.
 4
 5      9.5.4.5 Canopy Structure
 6           Another factor important in either O3 exposure or uptake, is how canopy structure affects
 7      O3 concentration in and under forest canopies. There have been several comprehensive studies
 8      of O3 concentrations under tree canopies (Enders, 1992; Fontan et al., 1992; Fredericksen et al.,
 9      1995; Joss and Graber, 1996; Kolb et al., 1997; Lorenzini and Nali, 1995; Neufeld et al., 1992;
10      Samuelson and Kelly, 1997).  In general they indicate a reduction in O3 of-20 to 40% below the
11      canopy but above the shrub/herb layers.  An essential component in the critical level AOT40 is
12      the height of where the O3 concentration is measured.  The critical levels are related to the O3
13      concentration measured at the top of the canopy, i.e., upper surface boundary of the  (quasi-)
14      laminar layer (Grunhage et al., 2003).  This location is presumably more related to stomatal
15      uptake. Weighting the O3 concentration at this location takes into account stomatal opening and,
16      if weighted with the Jarvis-Steward factors for radiation, temperature, and soil moisture, the
17      "lexicologically" effective AOT40 is obtained (Grunhage et al., 2003).
18           In a study that considers those factors important in O3 uptake that are also spatially
19      distributed as a result of canopy structure, Davison et al. (2003) reported that the variation in
20      visible injury in coneflower populations is unlikely to be due to differences in O3 flux and more
21      likely due to variation in PAR. At a height of 50 cm above ground, PAR was reduced by almost
22      90%, whereas the O3 varied from about 15 to 90% of ambient.  Ozone injury was not solely
23      related to O3 flux.  Although there have been studies of the effects of different light levels on O3
24      response, there have been few at the very low levels that occur in canopies of tall herbaceous
25      stands or in the ground layer of forests. Davison et  al. (2003) report that conductance was not
26      related to diurnal changes in light. The O3 levels were still about 90% of the O3 concentration
27      above the canopy when light was less than 5%. Light intensity dropped to 1.5% of open at 130
28      cm from the edge of the canopy, while O3 dropped to only 42%. The study, although reporting
29      on the adequacy of visible foliar injury as an indicator of O3 effects,  does suggest that
30      consideration of other factors such as light are important in predicting response. How this may
31      be included in developing exposure-response indices was not considered.

        January 2005                              9-173         DRAFT-DO  NOT QUOTE OR CITE

-------
 1      9.5.4.6  Site and Climate Factors
 2           Soil moisture is a critical factor in controlling O3 uptake through it's effect on plant water
 3      status and stomatal conductance.. In an attempt to relate uptake, soil moisture, and ambient air
 4      quality to identify areas of potential risk, available O3 monitoring data for 1983 to 1990 were
 5      used, along with literature-based seedling exposure-response data from regions within the
 6      southern Appalachian Mountains that might have experienced O3 exposures sufficient to inhibit
 7      growth (Lefohn et al., 1997).  In a small number of areas within the region, O3 exposures and
 8      soil moisture availability were sufficient to possibly result in growth losses in some sensitive
 9      species (e.g., black cherry). The conclusions were limited because of the interpolation of the O3
10      exposures in many of the areas and the hydrologic index used might not reflect actual water
11      stress.
12
13      9.5.4.7  Plant Defense Mechanism - Detoxification
14           The non-stomatal component of plant defenses are the most difficult to quantify, but some
15      studies are available (Barnes et al., 2002; Plochl et al., 2000; Chen et al., 1998; Massman and
16      Grantz, 1995). Massman et al. (2000) developed a conceptual model of a dose-based index to
17      determine plant injury response to O3 that relates to the traditional exposure-based parameters.
18      The index uses time-varying-weighted fluxes to account for the fact that flux is not necessarily
19      correlated with plant injury or damage.  Their model applies to plant foliar injury, and they
20      suggest that application of flux-based models for determining plant damage (yield or biomass)
21      will require a better understanding and quantification of the injury and damage relationship.
22
23      9.5.5 Ozone Uptake or Effective Dose as an Index
24           Developing an index that relates growth response to ambient exposure has been
25      approached in the past through various weighting functions on those ambient exposure factors,
26      including concentration, duration, and time of day for cumulating exposure. These indices have
27      not incorporated factors of climate patterns and species- and site factors that control O3 uptake
28      via canopy and stomatal conductance.  The other approach is basing the index on the O3
29      concentration going  into the leaf, or flux.  This approach includes those factors controlling
30      uptake through canopy and stomatal conductance and, by necessity, relies on models to predict
31      flux and ultimately the "effective" flux: that concentration that reaches the target tissue to cause

        January 2005                             9-174        DRAFT-DO NOT QUOTE OR CITE

-------
 1      an effect (Grunhage et al., 2004; Massman 2004; Massman et al., 2000) .Effective flux has been
 2      defined as the balance between the O3 flux and the detoxification process (Dammgen et al.,
 3      1993; Grunhage and Haenel, 1997; Musselman and Massman, 1999).  The time-integrated
 4      effective flux is termed "effective dose". The uptake mechanisms and the resistances in this
 5      process, including stomatal conductance and biochemical defense mechanisms, are discussed in
 6      the previous Section 9.3.
 7
 8      9.5.5.1  Models of Stomatal Conductance
 9          Only a limited number of studies have measured O3 concentration or its reaction products
10      within the leaf. Otherwise the index of uptake is a modeled result that considers site, climatic,
11      meteorological, and species-specific (e.g., detoxification reactions) factors. Models of O3
12      conductance into plant tissue are  available (Grunhage and Jager, 2003; Emberson et al., 2000a,b;
13      Grunhage and Haenel, 1997; Grunhage et al., 1997; Massman, 1993; Wesely, 1989). The
14      European Monitoring and Evaluation Program (EMEP) has an O3 deposition model developed
15      for application across Europe in conjunction with the EMEP photochemical model as a tool for
16      the critical levels program.  The model has been developed to estimate vegetation type-specific
17      O3 deposition  and stomatal flux, calculated according to a standard 3-resistance formulation
18      incorporating atmospheric, boundary layer, and stomatal resistances (Emberson 2000b).  The
19      model uses a multiplicative algorithm of the stomatal conductance of O3 (Jarvis, 1976) and has
20      been parameterized for 10 European tree species, 7 agricultural species, and 1 type of semi -
21      natural vegetation. The model calculates conductance as a function of leaf phenology,
22      temperature, photosynthetic flux  density (PFD), vapor pressure deficit (VPD), and soil moisture
23      (SMD). The environmental variables are site-specific (or regionally-specific). The most
24      important factors limiting O3 with this model were vapor pressure deficit (VPD), soil moisture
25      deficit, and phenology (Emberson et al., 2000b).  These factors demonstrate the critical linkage
26      of high VPD and stomatal closure, which typically co-occurs with high O3 concentrations.
27          A number of recent model-based studies relating flux and plant growth response have
28      investigated several crop  and forest tree species (Karlsson et al., 2004; Karlsson et al., 2004;
29      Pleijel et al., 2004; Matyssek et al., 2004; Elvira et al., 2004;  Soja et al., 2004; Weiser and
30      Emberson, 2004; Touvinen et al., 2004; Altimir, et al., 2004;  Gerosa et al., 2004; Mikkelsen,
31      et al., 2004, Bassin, et al., 2004; Emberson et al., 2000).  The studies have used earlier  exposure

        January 2005                              9-175        DRAFT-DO NOT QUOTE OR CITE

-------
 1      experiments as well as explicitly designed field studies but, to date, no abundance of data that
 2      would provide the basis for a flux-based index.
 3           The complexity of using flux as an index of O3 exposure for response is shown in field
 4      studies that measured O3 flux into Norway spruce and cembran pine (Wieser et al., 2000).  They
 5      demonstrated that stomatal conductance was the main limiting factor for O3 uptake and showed
 6      the dependence of that measure on crown position, needle age, and altitude. Consideration of the
 7      role of climate illustrates the importance of a flux measure. Pleijel et al., (2000) reported the
 8      improved relationship of yield in spring and winter wheat grown in OTCs in many areas across
 9      Europe when it  was related to the cumulative stomatal O3 uptake during the grain-filling period.
10      Compared to the AOT40, the cumulative uptake index estimated larger yield losses in the
11      relatively humid parts of western and northern Europe, while smaller yield loss was estimated
12      for the dry summer climates in south and central Europe.
13           Danielsson et al. (2003) compared the ability of two different stomatal models to relate
14      grain yield in field grown spring wheat to cumulated O3 uptake and an exposure index of
15      AOT40,  and found that the cumulated O3 uptake determined with either model performed better
16      in relating exposure to yield than did the cumulative exposure index of AOT40.
17           Cumulative O3 uptake was modeled for three deciduous and 2 confierous species growing a
18      different sites and elevations and compared with exposure measure of AOT40 at these sites
19      (Matyssek et al., 2004). A general linearity  was demonstrated between the two measures of O3
20      exposure; and, at any given AOT40, there was a 25 ± 11% variation in CU. Although no
21      correlation of growth alterations was observed with either the exposure or the uptake measure,
22      the modeled cumulative uptake was able to describe the variation in tree size and site location
23      which makes for a better measure in risk assessment of O3 (Matyssek et al., 2004). Karlsson
24      et al.  (2004) compared the biomass-response relationship in young trees at seven experimental
25      sites across Europe using modeled cumulative O3 uptake and AOT40. A weaker dose-response
26      relationships were reported for the cumulative uptake metric than the AOT40 (Karlsson et al.,
27      2004),
28           Concern about the complexity of the stomatal  models and the data needed to model O3
29      uptake, has led some researchers to offer modified accumulated exposure indices that consider
30      those meteorological factors controlling uptake (Karlsson et al., 2004; Gerosa et al., 2004). In a
31      study of  subterranean clover in Austria, Belgium, and southern Sweden, Karlsson et al., (2004)

        January 2005                             9-176       DRAFT-DO NOT QUOTE OR CITE

-------
 1      reported on the performance of a modified accumulated exposure over the threshold (mAOT)
 2      which was based on solar radiation and vapor pressure deficit.  This index improved the
 3      relationship for observed visible injury.  But if modeled uptake of O3 was derived from a simple
 4      stomatal conductance model considering solar radiation, VPD, and air temperature, then this
 5      index gave an even greater improvement in the relationship to visible injury than did the ambient
 6      exposure index of AOT40 (Karlsson et al., 2004B). The added value of the mAOT was
 7      worthwhile and it had a lower degree of complexity and data requirements modeling O3 uptake
 8      with stomatal models. Based on a study of O3 fluxes over a barley field in Italy, a similar
 9      modified exposure index was reported and referred to as "effective exposure" (Gerosa et al.,
10      2004). Their approach was similar in its consideration of physiological aspects in conjunction
11      with monitored O3 concentrations. It addressed the shortcoming of the data needs for modeled
12      O3 uptake.
13          Models that partition O3 uptake into stomatal and non-stomatal components are also now
14      available and predict a significant non-stomatal component in calculating O3 flux (Altimir et al.,
15      2004; Mikkelsen et al., 2004; Nikolov and Zeller, 2003; Zeller and Nikolov, 2000; Bassin et al.,
16      2004; Nussbaum et al., 2003). Altimir et al. (2004) compared the relative contribution of
17      stomatal and non-stomatal sinks at the shoot level for Scots pine.  Using the EMEP model with a
18      revised parameterization for Scots pine, they demonstrated that a major removal of O3 was due
19      to the non-stomatal component; and when a non-stomatal term was introduced dependent on
20      ambient relative humidity, the non-stomatal contribution to the total conductance was about
21      50%. Zeller and Nikolov (2000) demonstrated a large non-stomatal O3 uptake (41% of the total
22      annual flux) in subalpine fir at a site in southern Wyoming using the biophysical  model
23      FORFLUX. In a 5 year  study of measured O3 flux to a Norway spruce canopy, Mikkelsen et al.
24      showed monthly patterns of non-stomatal and stomatal deposition as part of total deposition to
25      the canopy. Their study demonstrated that daily means of O3 concentration and fluxes averaged
26      over 5 years correlate well, but the correlation is based on two different uncoupled processes
27      outside and inside the stomates. The destruction of O3 in the canopy outside the stomates is
28      influenced by temperature, light and humidity, e.g., surface reactions, NO- and VOC-emissions;
29      and these  same factors influence stomatal opening, e.g., mid-day and night closure.
30      Consequently, the diurnal O3 concentration and O3 flux do not correlate at all during the growing
31      season. The study estimated yearly stomatal uptake to be a minimum of 21% of total  deposition

        January 2005                             9-177       DRAFT-DO NOT QUOTE OR CITE

-------
 1      (i.e., non-stomatal as high as 80% of total). The stomatal uptake is highest in May-August (30-
 2      33%) and lowest in November-February (4-9%).
 3
 4      9.5.5.2 Nonlinear Response and Developing Flux Indices
 5           If only O3 flux were used as a metric to predict vegetation injury or damage, the prediction
 6      might be overestimated because of reported non-linear relationships between O3 and plant
 7      response (Amiro et al., 1984; Amiro and Gillespie, 1985; Bennett, 1979; U.S. Environmental
 8      Protection Agency, 1978; 1986; 1996). The non-linearity in the response surface suggested the
 9      existence of a biochemical threshold. More recently, nonlinear relationships between O3 flux
10      and yield were shown for potato (Pleijel et  al.,2002) and spring wheat (Danielsson et al.,2003).
11      The relationship between O3 flux and potato yield suggested using an instantaneous flux
12      threshold to overcome the nonlinear relationship (Pleijel et al., 2002). However, the authors did
13      not report a substantial improvement in the mathematical fitting of the model when applying the
14      threshold Most of the flux was accumulated below 0.06 ppm. Danielsson et al. (2003) was able
15      to show an improved relationship between O3 uptake and yield of spring wheat using a threshold
16      of 5 nmoles m"2 sec"1 (0.24 mg m"2 sec"1). These results suggest not all O3 entering the stomata
17      contribute to a reduction in yield, which depends to some degree on the amount of internal
18      detoxification occurring with each particular species (see Section 9.3).  The fact that the defense
19      and repair mechanisms vary diurnally as well as seasonally may make it extremely difficult to
20      apply a mathematically determined threshold to instantaneous flux measurements to  calculate
21      cumulative flux. The threshold models do not allow for the temporal (i.e., daily and  seasonal)
22      variability of defense mechanisms.
23           Musselman and Massman (1999) suggest that those species having high amounts of
24      detoxification potential may show less of a relationship of O3 stomatal uptake to plant response.
25      There has been no direct experimental demonstration of this relationship, however.  The cellular
26      detoxification reactions and repair processes which both detoxify oxidants as well as play central
27      role in the carbon economy of the plant are another level of resistance to O3 reaching the target
28      tissue (see Section 9.3).  As indicated earlier, effects occur on vegetation when the amount of
29      pollutant absorbed exceeds the ability of the plant to detoxify O3 or to repair the initial impact.
30      The magnitude of the response is determined by the amount of the pollutant reaching the target
31      site and the  ability of the  plant to reestablish homeostatic equilibrium. Thus, one would expect

        January 2005                             9-178       DRAFT-DO NOT QUOTE  OR CITE

-------
 1      to observe a decoupling of O3 uptake with vegetation effects, which would manifest itself as a
 2      nonlinear relationship between O3 flux and injury or damage.
 3           Additional factors for inclusion in flux-based models to predict vegetation effects are the
 4      defense and repair mechanisms.  Specifically, the relationship between conductance, O3
 5      concentration, and defense/repair mechanisms needs to be included. Recently, Massman (2004)
 6      described results that illustrate that the combination of conductance, O3 concentration, and
 7      diurnal variation of defense mechanisms show the daily maximum potential for plant injury
 8      (based on effective dose) coincides with the daily peak in O3 mixing ratio. Massman et al.
 9      (2000) stress that the product of the overlapping mathematical relationships of conductance,
10      concentration, and defense mechanisms results in a much different picture of potential impact to
11      vegetation than just the use of conductance and concentration in predicting vegetation effects.
12
13      9.5.5.3  Simulation Models
14           Another approach for determining O3 uptake and relating growth response to ambient O3
15      exposure may be the use of physiologically-based simulation models.  Several of these have
16      been used in various contexts, comparing O3 response in a number of tree species with varying
17      climate and site factors (e.g., soil moisture) (e.g., Laurence et al., 2001; Ollinger et al., 1997;
18      Ollinger et al., 1998; Weinstein et al., 2001; Weinstein et al., 2004; Tingey et al., 2000; Tingey
19      et al., 2004). One of the important considerations in applying simulation modeling is to
20      carefully assess the uncertainties associated with the modeling predictions.  Further efforts need
21      to be made to exercise the models so that they predict past growth losses  associated with changes
22      in O3 exposures that can be verified with on-the-ground surveys.
23
24      9.5.6  Summary
25           A few studies published since 1996 have substantiated earlier conclusions on the role of O3
26      exposure components (including concentration, duration and exposure patterns) in describing
27      growth response to O3 exposures. Recent studies of different exposure patterns have confirmed
28      earlier studies on the importance of higher concentrations and duration of exposure when
29      describing response. An inferred role of peak concentrations is possible from consideration of
30      improved air quality in regions like  the San Bernardino Mountains in southern California.
        January 2005                             9-179        DRAFT-DO NOT QUOTE OR CITE

-------
 1      Studies provide the basis for focusing on the higher O3 concentrations, while including the lower
 2      levels, when estimating the effects of emission reductions on vegetation.
 3           A few studies have demonstrated the potential disconnection of peak events and maximal
 4      stomatal conductance.  In addition, a few studies have demonstrated the uptake of O3 during
 5      nighttime hours and suggested the need to cumulate O3 exposure 24 h per day and not just during
 6      daylight hours.
 7           Several studies since 1996 have demonstrated another critical concern in developing an
 8      index for exposure. The concern is that peak O3 events and maximum stomatal conductance
 9      may be temporally separate. This disconnection introduces uncertainty in assessing O3 impact
10      when using the current ambient exposure based cumulative, concentration weighted indices.
11      If stomatal conductance is relatively low, as in the late afternoon in arid climates, and that is the
12      same time as the peak O3 concentrations, then use of an exposure index that does not consider
13      this disconnect, will overestimate the effect of the exposure. This concern is especially apparent
14      when assessing the impact of O3 across all the varied climatic regions of the United States or
15      Europe.  Some studies use stomatal models to predict uptake or using process level models (e.g.,
16      TREGRO) to integrate those climate and site factors that drive this temporal pattern of stomatal
17      conductance and exposure, and thus reduce some of the uncertainty in regional  to national
18      assessments of effects.  These approaches however are still species dependent.
19           The results of these studies and reviews have indicated the  need to continue to develop
20      indices that are more physiologically and meteorologically  connected to the actual dose of O3 the
21      plant receives.  The cumulative concentration-weighted exposure indices are acknowledged
22      surrogates for effective dose that are simple conceptually and easy to measure.  They do not fully
23      characterize the potential for plant uptake and resulting effects associated with O3 because the
24      indices, being measures of ambient concentration, do not include the physical, biological, and
25      meteorological processes controlling the transfer of O3 from the atmosphere through the leaf and
26      into the leaf interior. Use of such indices is especially limited in spatial risk characterizations
27      because of the lack of linkage of meteorology, species and site-specific factors influencing O3
28      uptake.  The flux-based approach should provide an opportunity  to improve upon the
29      concentration-based (i.e., exposure indices) approach. A cautionary argument was advanced in a
30      few publications centered around the non-linear relationship between ozone uptake and plant
31      injury (not growth alterations) response.  The concern was that not all O3 stomatal uptake results

        January 2005                             9-180        DRAFT-DO NOT QUOTE OR CITE

-------
 1      in a reduction in yield, which depends to some degree on the amount of internal detoxification
 2      occurring with each particular species.  Those species having high amounts of detoxification
 3      potential may show less of a relationship between O3 stomatal uptake and plant response.
 4           The European approach and acceptance of flux-based critical values is a recognition of this
 5      problem; and a concerted research effort is needed to develop the necessary experimental data
 6      and modeling tools that will provide the scientific basis for such critical levels for O3 (Grunhage
 7      et al., 2004; Fuhrer et al., 1997; Grunhage and Jager, 1994).
 8           At this time, based on the current state of knowledge, exposure indices that differentially
 9      weight the higher hourly average O3 concentrations, but include the mid-level values, represent
10      the best approach in the United States for relating vegetation effects to O3 exposure..  A large
11      database exists that has been used for establishing exposure-response relationships.  Such a
12      database does not yet exist for relating O3 flux to growth response.  Those pattern disconnects
13      between period of uptake and peak occurrence, as well as the potential for nocturnal uptake,
14      should be considered with some weighting functions in the currently used exposure indices.
15      Of particular consideration would be their inclusion in regional to national estimations of O3
16      impacts on vegetation.  Another useful approach to regional assessment for given specie(s) is
17      simulating growth effects with process-based models that account for seasonal climate and site
18      factors that control conductance
19           It is anticipated that, as the overlapping mathematical relationships of conductance,
20      concentration, and defense mechanisms are better defined, O3 flux-based models may be able to
21      predict vegetation injury and/or damage at least for some categories of canopy-types with more
22      accuracy than the exposure-response models.
23
24
25      9.6  OZONE EXPOSURE-PLANT RESPONSE RELATIONSHIPS
26      9.6.1  Introduction
27           Ambient O3 concentrations have long been known to cause visible symptoms, decreases in
28      photosynthetic rates, decreases in plant growth, and decreases in the yield of marketable organs
29      (U.S. Environmental Protection Agency, 1978, 1986,  1996). Yet, despite considerable research
30      in the U.S. and other countries during the past three decades, quantifying the effects of ambient
31      O3 exposure on vegetation remains a challenge. Numerous studies have related O3 exposure to

        January 2005                             9-181        DRAFT-DO NOT QUOTE OR CITE

-------
 1      plant responses, with most effort focused on the yield of crops and the growth of tree seedlings.
 2      Most experiments exposed individual plants grown in pots or soil under controlled conditions to
 3      known concentrations of O3 for a segment of daylight hours for some portion of the plant's life
 4      span (Section 9.2). The response of a plant species or variety to O3 exposure depends upon
 5      many factors discussed in previous sections, including genetic characteristics (Section 9.4.2),
 6      biochemical and physiological status (Section 9.4), and previous and current exposure to other
 7      stressors (Sections 9.4, 9.5).  Section 9.4 describes how O3 moves from the atmosphere into the
 8      leaf and the subsequent biochemical and physiological responses of plants. The current section
 9      focuses on the quantitative responses of plants to seasonal or multiyear exposures  to known
10      amounts of O3. Quantitative responses include foliar symptoms and decreased growth of whole
11      plants or decreased harvestable portions of them.  Because of the available information, most of
12      this section focuses on the response of individual plants, especially crop plants and tree
13      seedlings, with limited discussion of mixtures of herbaceous species.  The responses of natural
14      ecosystems are discussed in Section 9.7.
15           This section will pay particular attention to  studies conducted since the publication of the
16      1996 AQCD (U.S. Environmental Protection Agency, 1996).  However, because much O3
17      research was conducted prior to the 1996 AQCD, the present discussion of vegetation response
18      to O3 exposure is largely based on the conclusions of the 1978, 1986,  and 1996 criteria
19      documents (U.S. Environmental Protection Agency, 1978, 1986, 1996). To provide a context for
20      the discussion of recent research,  the the key findings and conclusions of those three documents
21      are first summarized below.
22
23      9.6.2  Summary of Key Findings/Conclusions from  Previous Criteria
24             Documents
25           Experimental data reviewed in the 1978 and 1986 criteria documents dealt primarily with
26      the effects of O3 on agricultural crop species (U.S. Environmental Protection Agency, 1978,
27      1986). The chapter on vegetation effects in the 1978 document (U.S. Environmental Protection
28      Agency, 1978) emphasized foliar symptoms and growth effects, but not those effects that
29      affected yield, an emphasis dictated by the kind of data available at the time. The  1986
30      document reviewed a substantial new body of evidence based on open-top chamber (OTC)
31      experiments (see Section 9.2) showing  that ambient O3 exposures reduced the growth and yield

        January 2005                             9-182       DRAFT-DO NOT QUOTE OR CITE

-------
 1      of herbaceous plants, again with a focus on major crop species.  In the 1986 and 1996
 2      documents, data were presented from regression studies conducted to develop exposure-response
 3      functions for estimating yield loss of major crop species in different regions of the United States.
 4      The 1996 document included results from additional herbaceous crop species as well as shrub
 5      and tree species.  For a number of tree species, biomass growth of seedlings was related to
 6      growing season O3 exposures to produce response functions for estimating O3 exposures that
 7      reduce growth by 10 or 30%. Also, in the 1986 and 1996 documents, data from studies using
 8      ethylene diurea (EDU) as a protectant were reviewed. The 1978, 1986, and 1996  criteria
 9      documents also reviewed data on the response to O3 exposures of forest ecosystems in the
10      San Bernardino Mountains of southern California (U.S. Environmental Protection Agency, 1978,
11      1986,  1996). Because this region is exposed to high concentrations of O3  and has shown
12      evidence of ecosystem-level changes, it remains an important study area (see Section 9.7).
13           Ozone can cause a range of effects, beginning with individual cells,  leaves, and plants, and
14      proceeding to plant populations and communities. These effects may be classified as either
15      injury or damage. Injury encompasses all plant reactions, such as reversible changes in plant
16      metabolism (e.g., altered photo synthetic rate), altered plant quality, or reduced growth that does
17      not impair yield or the intended use or value of the plant (Guderian, 1977). In contrast, damage
18      includes all effects that reduce or impair the intended use or value of the plant.  Damage includes
19      reductions in aesthetic values as well as losses in terms of weight, number, or size of the plant
20      part that is harvested (yield loss).  Yield loss also may include changes in  crop quality, i.e.,
21      physical appearance, chemical composition, or the ability to withstand storage.  Losses in
22      aesthetic values are difficult to quantify. Although foliar symptoms cannot always be classified
23      as damage, their occurrence indicates that phytotoxic concentrations of O3 are present, and,
24      therefore, studies should be conducted to assess the  risk to vegetation.
25           Visible symptoms due to O3 exposures reduce the market value  of certain crops and
26      ornamentals for which leaves are the product, e.g., spinach, petunia, geranium, and poinsettia.
27      The concept of limiting values used to summarize foliar symptoms in the  1978 document (U.S.
28      Environmental Protection Agency, 1978) was also considered valid in the 1986 document (U.S.
29      Environmental Protection Agency, 1986). Jacobson (1977) developed limiting  values by
30      assessing the available scientific literature and identifying the lowest exposure
31      concentration/duration reported to cause foliar symptoms in a variety of plant species.

        January 2005                              9-183        DRAFT-DO NOT QUOTE OR CITE

-------
 1      A graphical analysis presented in those documents indicated that the limit for reduced plant
 2      performance was an exposure to 50 ppb for several hours per day for more than 16 days.
 3      Decreasing the exposure period to 10 days increased the concentration required to cause
 4      symptoms to 100 ppb; and a short, 6-day exposure further increased the concentration to cause
 5      symptoms to 300 ppb.  These limiting values established in 1978 were still deemed appropriate
 6      in the 1986 and 1996 criteria documents.  Such foliar symptoms are caused by O3 concentrations
 7      that occur in the United States as shown in Table 9-13 (as adapted from U.S. Environmental
 8      Protection Agency, 1996).
 9           The 1986 document emphasized that, although foliar symptoms on vegetation are often an
10      early and obvious manifestation of O3 exposure, O3 effects are not limited to foliar symptoms.
11      Other effects include reduced growth of many organs (including roots), changes in crop quality,
12      and alterations in plant susceptibility to biotic stressors and sensitivity to abiotic stressors. The
13      1986 document also emphasized that O3 exerts phytotoxic effects only if a sufficient amount of
14      O3 reaches sensitive sites within the leaf (Section 9.3).  Ozone injury will not occur if the rate of
15      O3 uptake is low enough that the plant can detoxify or metabolize O3 or its metabolites or if the
16      plant is able to repair or compensate for the effects (Tingey and Taylor, Jr., 1982; U.S.
17      Environmental Protection Agency, 1986).  Cellular disturbances that are not repaired or
18      compensated for are ultimately expressed as foliar symptoms, reduced root growth, or reduced
19      yield of fruits or seeds.
20           Beginning in the  1986 document and continuing in the 1996 document, OTC studies were
21      reviewed that better quantified the relationship between O3  exposure and effects on crop species,
22      with a focus on yield loss. These studies can be grouped into two types, depending on the
23      experimental design and statistical methods used to analyze the data:  (1) studies that developed
24      predictive equations relating O3 exposure to plant response, and (2) studies that compared the
25      effects of discrete treatment level(s) to a control. The advantage  of the regression approach is
26      that exposure-response models can be used to interpolate results between treatment levels.
27           Discrete treatment experiments were designed to test whether specific O3 treatments were
28      different from the control rather than to develop exposure-response equations, and the data were
29      analyzed using analyses of variance.  When summarizing these studies using discrete treatment
30      levels, the lowest O3 concentration that significantly  reduced yield was determined from
31      analyses done by the original authors.  Often, the lowest concentration used in the study was the

        January 2005                             9-184        DRAFT-DO NOT QUOTE OR CITE

-------
       Table 9-13. Summary of Ozone Exposure Indices Calculated for 3- or 5-Month
                               Growing Seasons from 1982 to 1991a
3-month growing season (June-August)
No of
Year Sites"
1982 99
1983 102
1984 104
1985 117
1986 123
1987 121
1988 139
1989 171
1990 188
1991 199
Among Years
HDM2C
ppm
Mean CVd
0.114 23.7%
0.125 24.9%
0.117 24.6%
0.117 24.6%
0.115 21.8%
0.119 22.9%
0.129 21.3%
0.105 23.1%
0.105 21.6%
0.106 22.0%
0.113 11.1%
M7
ppm
Mean
0
0
0
0
0
0
0
0
0
0
0
CV
18.7%
21.9%
18.2%
17.1%
19.1%
17.6%
17.8%
17.5%
18.3%
18.4%
10.0%
SUMOO
ppm-h
Mean
82.9
86.1
84.1
84.6
85.3
86.9
97.6
86.4
85.7
87.7
87.0
CV
19.1%
22.1%
19.9%
18.0%
18.0%
17.3%
19.6%
19.9%
21.0%
21.3%
9.9%
SUM06
ppm-h
Mean CV
26.8 68.8%
34.5 58.1%
27.7 58.4%
27.4 59.6%
27.7 65.0%
31.2 56.4%
45.2 46.8%
24.8 78.7%
25.8 76.2%
28.3 74.2%
29.5 42.1%
SIGMOID
ppm-h
Mean
26.3
33.0
27.4
27.4
27.7
30.4
42.9
25.8
26.6
28.9
29.4
CV
56.7%
52.3%
47.9%
47.6%
51.8%
46.8%
42.4%
59.4%
59.2%
59.5%
31.0%
5-month growing season (May-September)
No. of
Year Sites
1982 88
1983 87
1984 95
1985 114
1986 118
1987 116
1988 134
1989 158
1990 172
1991 190
Among Years
M7
ppm
Mean
0.048
0.051
0.048
0.048
0.048
0.050
0.054
0.047
0.049
0.050
0.049
SUMOO
ppm-h
cv
20.6%
22.1%
18.0%
18.4%
20.3%
20.3%
18.7%
18.6%
19.8%
19.8%
9.8%
Mean
122.9
129.6
126.2
124.5
123.3
128.7
141.7
127.8
129.4
130.6
129.0
CV
22.3%
24.4%
19.1%
19.4%
21.4%
20.4%
22.0%
22.5%
22.7%
23.6%
9.9%
SUM06
ppm-h
Mean
37.3
44.4
36.7
36.2
34.9
42.2
58.0
32.7
34.6
36.8
38.7
CV
70.9%
61.9%
60.8%
63.8%
70.7%
62.0%
50.5%
87.8%
82.7%
80.7%
42.5%
SIGMOID
ppm-h
Mean
37.1
43.8
37.6
37.0
35.6
41.8
55.6
35.2
37.0
38.8
39.6
CV
57.8%
52.7%
46.9%
50.3%
55.7%
50.3%
45.0%
64.1%
62.1%
62.9%
29.8%
 "Updated and additional years from data given in Table III of Tingey et al. (1991), where the spatial and temporal variation
  in ambient O3 exposures is expressed in terms of several exposure indices.
 b Indicates the number of separate monitoring sites included in the analysis; fewer sites had 5 months of available data than
  had 3 months of available data.
 c The 2HDM index is calculated for sites with at least 3 months of available data.  SUMOO, SUM06, M7, SIGMOID, and
  2HDM are the cumulative sum above 0.0 ppm, the cumulative sum above 0.06 ppm, the 7-h seasonal mean, the sigmoid
  weighted summed concentration, and the second highest daily maximum 1-h concentration, respectively.
 d C V = coefficient of variation.

 Source:  Table 5-30 from U.S. Environmental Protection Agency (1996) based on Tingey etal. (1991).
January 2005
9-185
DRAFT-DO NOT QUOTE OR CITE

-------
 1      lowest concentration reported to reduce yield; hence, it was not always possible to estimate a no-
 2      effect exposure concentration.  In general, the data indicated that 100 ppb O3 (frequently the
 3      lowest concentration used in the studies) for a few hours per day for several days to several
 4      weeks usually caused significant yield reductions of 10 to 50%.
 5           By the time the 1986 document was prepared, much new information concerning the
 6      effects of O3 on the yield of crop plants had become available through EPA's NCLAN research
 7      program and through research funded by other agencies. The NCLAN project was initiated by
 8      EPA in 1980 primarily to improve estimates of yield loss under field conditions and to estimate
 9      the magnitude of crop losses caused by O3 throughout the United States (Heck et al., 1982,
10      1991). The cultural conditions used in the NCLAN studies approximated typical agronomic
11      practices. The primary objectives were:
12            (1)   to define relationships between yields of major agricultural crops and O3 exposure
                   as required to provide data necessary for economic assessments and development
                   ofO3NAAQS;
13            (2)   to assess the national economic consequences resulting from O3 exposure of major
                   agricultural crops; and
14            (3)   to advance understanding of cause-and-effect relationships that determine crop
                   responses to pollutant exposures.
15           Using NCLAN data, the O3 concentrations predicted to cause 10 or 30% yield loss were
16      estimated using linear or Weibull response functions. The data in Table 9-14 are from the 1996
17      document and were based on yield-response functions for 38 species or cultivars developed from
18      studies using OTCs of the type developed by Heagle et al. (1973) (see Section 9.2).  Composite
19      exposure-response functions for both crops and tree  seedlings as a function of O3 exposure
20      expressed as SUM06 are shown in Figure 9-18. Review of these data indicate that 10% yield
21      reductions could be predicted for more than 50% of experimental cases when:  (1) 12-h SUM06
22      values exceeded 24.4 ppm-h, (2) SIGMOID values exceeded 21.5 ppm-h,  or (3) 7-h seasonal
23      mean concentrations were 50 ppb.  The SIGMOID index is very similar to the W126 index (see
24      Section 9.5 for further information about O3 indices). Much lower values are required for each
25      index to protect 75% of experimental cases (Table 9-15).  Grain crops were generally found to
26      be less sensitive than other crops.  The data summarized in the 1996 criteria document also
27      indicated that the variation in sensitivity within species may be as great as differences between
28      species.

        January 2005                             9-186       DRAFT-DO NOT QUOTE OR CITE

-------
     Table 9-14. Ozone Exposure Levels (Using Various Indices) Estimated To Cause at
                  Least 10% Crop Loss in 50 and 75% of Experimental Cases"
50th PERCENTILE"
NCLAN Data (n = 49; wet and dry)d
NCLAN Data (n = 39; wet only)
NCLAN Data (n = 54; wet and dry)6
NCLAN Data (n = 42; wet only)6
NCLAN Data (n = 10; wet)
NCLAN Data (n = 10; dry)
Cotton Data (n = 5)
Soybean Data (n = 13)
Wheat Data (n= 6)
Cotton Data (n = 5)e
Soybean Data (n=15)e
Wheat Data (n = 7)e
SUM06
24.4
22.3
26.4
23.4
25.9
45.7
23.6
26.2
21.3
30.0
23.9
25.9
SEC
3.4
1.0
3.2
3.1
4.5
23.3
2.3
5.4
15.2
12.7
6.5
10.5
SIGMOID
21.5
19.4
23.5
22.9
23.4
40.6
19.3
22.6
19.3
27.2
22.0
21.4
SE
2.0
2.3
2.4
4.7
3.2
0.1
2.3
3.6
12.7
12.8
8.0
9.4
M7
0.049
0.046
NA
NA
0.041
0.059
0.041
0.044
0.061
NA
NA
NA
SE
0.003
0.003
NA
NA
0.001
0.014
0.001
0.005
0.018
NA
NA
NA
2HDM
0.094
0.090
0.099
0.089
0.110
0.119
0.066
0.085
0.098
0.075
0.088
0.097
SE
0.006
0.010
0.011
0.008
0.042
0.017
0.032
0.013
0.059
0.012
0.008
0.028
75th PERCENTILE"
NCLAN Data (n = 49; wet and dry)
NCLAN Data (n = 39; wet only)
NCLAN Data (n = 54; wet and dry)6
NCLAN Data (n = 42; wet only)6
NCLAN Data (n = 10; wet)
NCLAN Data (n = 10; dry)
Cotton Data (n = 5)
Soybean Data (n = 13)
Wheat Data (n = 6)
Cotton Data (n = 5)e
Soybean Data (n=15)e
Wheat Data (n = 7)e
14.2
14.3
16.5
17.2
16.4
24.0
21.8
14.2
11.7
21.1
15.3
5.1
4.2
2.7
4.3
3.0
3.7
0.8
5.0
0.1
2.5
6.0
4.1
2.6
11.9
12.6
14.5
14.7
13.7
22.3
17.5
12.4
10.9
16.7
13.4
8.5
5.6
2.3
3.2
2.4
3.2
0.1
2.8
0.1
2.4
5.7
4.1
3.4
0.040
0.039
NA
NA
0.040
0.053
0.041
0.041
0.054
NA
NA
NA
0.007
0.005
NA
NA
0.001
0.022
0.001
0.006
0.032
NA
NA
NA
0.051
0.056
0.073
0.070
0.080
0.093
0.065
0.069
0.062
0.070
0.078
0.054
0.010
0.006
0.006
0.006
0.032
0.003
0.014
0.004
0.035
0.034
0.007
0.027
 aSee Appendix A for abbreviations and acronyms.
 bThe numbers in parentheses are the number of cases used in deriving the various exposure levels.
 °Standard error (SE).
 dNCLAN data refers to studies conducted as part of the NCLAN project. Wet and dry refer to watering regimes
  used in the studies, wet being well-watered, and dry meaning some level of drought stress was imposed.
 e24-h exposure statistics reported in Lee et al. (1994b). Relative yield loss for 2HDM is relative to yield at 40 ppb
  rather than 0 ppb as was used in Tingey et al. (1991).

 Source:  U.S. Environmental Protection Agency (1996), (modified from Tingey et al. (1991).
January 2005
9-187
DRAFT-DO NOT QUOTE OR CITE

-------
              o
              '•B
100%

90%

80%

70%

60%

50%

40%

30%

20%

10%

 0%
                      a. Crops
                                                                 75th Percentile
                                                                 50th Percentile
                                                                 25th Percentile
                         \
                        10
                I
               20
 I
30
 I
40
 I
50
                                                             60
                                   24-h SUM06 (ppm-h)
^_^
«"
o
w
03
E
o
5
•a
•2
1
£
100%-q
90% -1
80% -|
70% -I
60% -I
50% -1
40% -1
30% -|
20% -i
10%-|
n% -
b. Tree Seedlings




75th Percentile

	 	 50th Percentile
	 	 	 	 ~~ 25th Percentile
                         \
                        10
                \
               20
 \
30
 \
40
                                                      50      60
                          24-h SUM06 (ppm-h) (adjusted to 92 days)
Figure 9-18.  Distribution of biomass loss predictions from Weibull and linear exposure-
             response models that relate biomass to O3 exposure. Exposure is
             characterized with the 24-h SUM06 statistic using data from (a) 31 crop
             studies from National Crop Loss Assessment Network (NCLAN) and
             (b) 26 tree seedling studies conducted at U.S. Environmental Protection
             Agencys' Environmental Research Laboratory in Corvallis, OR; Smoky
             Mountains National Park, TN; Houghton, Michigan; and Delaware, Ohio.
             Separate regressions were calculated for studies with multiple harvests or
             cultivars, resulting in a total of 54 individual equations from the 31 NCLAN
             studies and  56 equations from the 26 seedling studies. Each equation was
             used to calculate the predicted relative yield or biomass loss at 10, 20, 30, 40,
             50, and 60 ppm-h, and the distributions of the resulting loss were plotted.
             The solid line is the calculated Weibull fit at the 50th percentile.

Source: U.S. Environmental Protection Agency (1996); Hogsett et al. (1995).
January 2005
                        9-188
             DRAFT-DO NOT QUOTE OR CITE

-------
 1           The chemical protectant, ethylene diurea (EDU), was also used to provide estimates of
 2     yield loss.  The impact of O3 on yield was determined by comparing the yield data from plots
 3     treated with EDU versus untreated plots.  Studies indicated that yields were reduced by 18 to
 4     41% when daytime ambient O3 concentrations exceeded 80 ppb for 5 to 18 days over the
 5     growing season. For this approach to be credible, the effects of EDU itself on a particular
 6     species must be preestablished under conditions without O3 exposure (Kostka-Rick and
 7     Manning, 1992).
 8           The 1996 criteria document reviewed several experiments demonstrating that the seedlings
 9     of some tree species such as poplars  (Populus) and black cherry are as sensitive to O3 as are
10     annual plants, in spite of the fact that trees are longer-lived and generally have lower rates of gas
11     exchange, and,  therefore, a lower uptake of O3.  The 1996 document also reviewed data showing
12     that O3 exposures that occur at present in the United States are sufficient to affect the growth of a
13     number of trees species.  For example, exposure-response functions for 51 cases of tree seedling
14     responses to O3, including 11 species representing deciduous and evergreen growth habits,
15     suggest that a SUM06 exposure for 5 months  of 31.5 ppm-h would protect hardwoods from a
16     10% growth loss in 50% of the cases studied (Table 9-15). Similarly, a SUM06 exposure of
17     42.6 ppm-h should provide the same  level of protection for evergreen seedlings. However, these
18     results do not take into the account the possibility of effects on growth in subsequent years.
19     Because multiple-year exposures may cause a cumulative effect on the growth of some trees
20     (Simini et al., 1992; Temple et al., 1992), it is likely that a number of species are  currently being
21     affected even at ambient exposures (Table 9-14).
22           In 1986, the EPA (U.S. Environmental Protection Agency, 1986) established that 7-h per
23     day growing season mean exposures to O3 concentrations above 50 ppb were likely to cause
24     measurable yield loss in agricultural  crops.  At that time,  few conclusions could be drawn about
25     the response of deciduous or evergreen trees or shrubs, due to the lack of information about
26     response of such plants to season-long exposures to O3 concentrations of 40 to 60 ppb and above.
27     However, the 1978 and 1986 criteria documents (U.S. Environmental Protection Agency, 1978,
28     1986) indicated that the limiting value for foliar symptoms on trees and shrubs was 60 to 100
29     ppb for 4 h. From 1986 to 1996, extensive research was conducted, establishing the sensitivity
30     of many tree species.  Based on research published since  the 1986 criteria document,, a number
31     of conclusions were drawn in 1996 AQCD (U.S. Environmental Protection Agency, 1996):

       January 2005                              9-189        DRAFT-DO NOT QUOTE OR CITE

-------
      Table 9-15. SUM06 Levels Associated with 10 and 20% Total Biomass Loss for
                            50 and 75% of the Seedling Studies
         (The SUM06 value is adjusted to an exposure length of 92 days per year.)a

 Weibull Equations (all 51 seedling studies):

 50th Percentile PRYL1 = 1 - exp(-[SUM06/176.342]** 1.34962)
 75th Percentile PRYL = 1 - exp(-[SUM06/104.281]** 1.46719)


 Weibull Equations (27fast-growing seedling studies):

 50th Percentile PRYL = 1 - exp(-[SUM06/150.636]** 1.43220)
 75th Percentile PRYL = 1 - exp(-[SUM06/89.983]** 1.49261)


 Weibull Equations (24 slow to moderate growing seedling studies):

 50th Percentile PRYL = 1 - exp(-[SUM06/190.900]** 1.49986)
 75th Percentile PRYL = 1 - exp(-[SUM06/172.443]**l.14634)


 Weibull Equations (28 deciduous seedling studies):

 50th Percentile PRYL = 1 - exp(-[SUM06/142.709]** 1.48845)
 75th Percentile PRYL = 1 - exp(-[SUM06/87.724]** 1.53324)


 Weibull Equations (23 evergreen seedling studies):

 50th Percentile PRYL = 1 - exp(-[SUM06/262.911]** 1.23673)
 75th Percentile PRYL = 1 - exp(-[SUM06/201.372]** 1.01470)


             Levels Associated with Prevention of a 10 and 20% Total Biomass Loss
	for 50 and 75% of the Seedlings	

 All 51 Seedling Cases
                                            Percent of Seedlings
                                         50%                75%

        Relative            10%          33.3                22.5

        Biomass Loss       20%          58.0                37.5


 27 Fast-Growing Seedling Cases
                                           Percent of Seedlings

Relative
Biomass Loss

10%
20%
50%
31.3
52.9
75%
19.4
32.4
January 2005                               9-190       DRAFT-DO NOT QUOTE OR CITE

-------
  Table 9-15 (cont'd). SUM06 Levels Associated with 10 and 20% Total Biomass Loss for
                           50 and 75% of the Seedling Studies
        (The SUM06 value is adjusted to an exposure length of 92 days per year.)a

 24 Slow-to-Moderate-Growth Seedling Cases
                                          Percent of Seedlings

                                        50%               75%

        Relative           70%          42.6               24.2

        Biomass Loss       20%          70.2               46.6

 28 Deciduous Seedling Cases
                                          Percent of Seedlings

                                        50%               75%

        Relative           70%          31.5               20.2

        Biomass Loss       20%          52.1                33

 23 Evergreen Seedling Cases
                                          Percent of Seedlings
                                        50%              75%

        Relative          70%            42.6              21.9

        Biomass Loss      20%            78.2              45.9
 aSee Appendix [XXX] for abbreviations and acronyms.
 bPRYL = predicted relative yield (biomass) loss

 Source: U.S. Environmental Protection Agency (1996), based onHogsett et al. (1995).
       (1)   An analysis of 10 years of monitoring data from more than 80 to almost 200 non-
            urban sites in the United States established ambient 7-h growing season average
            concentrations of O3 for 3 or 5 months of 51 to 60 ppb and 47 to 54 ppb,
            respectively.  The SUM06 exposures ranged (a) from 24.8 to 45.2 ppm-h for
            3 months and (b) from 32.7 to 58.0 ppm-h for 5 months (Tingey et al. (1991),
            Table 9-14).

       (2)   The results of OTC studies that compared yields at ambient O3 exposures with
            those in filtered air and retrospective analyses of crop data (Table 9-14) established
            that ambient O3 concentrations were sufficient to reduce the yield of major crops in
            the United States. Research results since 1978 did not invalidate EPA conclusions
            (U.S. Environmental Protection Agency, 1978, 1986) that foliar symptoms due to
January 2005                             9-191        DRAFT-DO NOT QUOTE OR CITE

-------
                    O3 exposures reduce the market value of certain crops and ornamentals where
                    leaves are the product (such as spinach, petunia, geranium, and poinsettia) and that
                    such damage occurs at ambient O3 concentrations observed in the United States.

 3            (3)   A 3-month SUM06 exposure of 24.4 ppm-h, corresponding to a 7-h mean of 49 ppb
                    and  a 2HDM of 94 ppb O3 may prevent a 10% loss in 50% of the 49 experimental
                    cases analyzed by Tingey et al. (1991).  A 12-h growing season mean of 0.045 ppb
                    should restrict yield losses to 10% in major crop species (Lesser et al., 1990).

 4            (4)   Depending on duration, concentrations of O3 and SUM06 exposures currently in
                    the United States are sufficient to affect the growth of a number of tree species.
                    Given the fact that multiple-year exposures may cause a cumulative effect on the
                    growth of some trees (Simini et al., 1992; Temple et al., 1992), it is likely that a
                    number of species currently are being impacted, even at ambient O3 exposures
                    (Table 9-14).

 5            (5)   Exposure-response functions for 51 cases of seedling response to O3 (Hogsett et al.,
                    1995), including 11 species representing deciduous and evergreen growth habits,
                    suggest that a SUM06 exposure for 5 months of 31.5 ppm-h would protect
                    hardwoods from a 10% growth loss in 50% of the cases studied. A SUM06
                    exposure of 42.6 ppm-h should provide the same level of protection for evergreen
                    seedlings. Note that these conclusions do not take into the account the possibility
                    of effects on growth in subsequent years, an important consideration in the case of
                    long-lived species.

 6            (6)   Studies of the response of trees to O3 have established that, in some cases (for
                    instance, poplars and black cherry), trees are as sensitive to O3 as are annual plants,
                    in spite of the fact that trees are longer-lived and generally have lower gas
                    exchange rates, and, therefore, lower O3 uptake.

 7            (7)   Use of the chemical protectant, EDU, is of value in estimating O3-related losses in
                    crop yield and tree growth, provided that care is exercised in establishing
                    appropriate EDU dosages to protect the plants without affecting growth.

 8           The major question to be addressed in the remainder of this section is whether new

 9     information supports or alters the 1996 criteria document conclusions summarized above.

10     In particular, this  section  evaluates whether the response of plants to experimental treatments at

11     or near O3 concentrations characteristic of ambient levels in many areas of the United States

12     (Table 9-14) can be compared to a control or reduced O3 treatment to establish a potential

13     adverse effect.  Before evaluating new information from the literature on O3 effects on

14     vegetation, O3 exposure indices used in O3 studies and trends in O3 exposure patterns during the

15     past two decades  are briefly reviewed.
16


       January 2005                             9-192       DRAFT-DO NOT QUOTE OR CITE

-------
 1      9.6.3  Ozone Indices and Ambient Exposure
 2           As recognized in both the 1986 and the 1996 criteria documents, the characterization and
 3      representation of the exposure of vegetation to O3 is problematic, because the specific aspects of
 4      pollutant exposure that cause injury or damage are difficult to quantify. This issue is addressed
 5      in Section 9.5, and only a few points will be discussed here in order to provide a context for
 6      interpreting data on exposure-response relationships. The most important effects of O3 on
 7      vegetation occur due to uptake of O3 through stomata, with subsequent oxidative injury that
 8      appears to be rather nonspecific (Section 9.3). As has been discussed by numerous authors
 9      during the last three decades, from a toxicological and physiological view, it is much more
10      realistic to relate effects to internal (absorbed) O3 dose rather than to exposure near the leaf or
11      canopy (Fowler and Cape, 1982; Fuhrer et al., 1992; Griinhage et al., 1993, 1999; Legge et al.,
12      1995; Massman et al., 2000; Musselman and Massman, 1999; Pleijel et al., 1995; Runeckles,
13      1974; Taylor, Jr. et al., 1982; Tingey and Taylor, Jr., 1982) (see also Section 9.5).  Theoretically,
14      flux estimates should improve the assessment of O3 effects, but despite recent attention to this
15      topic, particularly in Europe, it remains difficult to estimate flux in the field outside of
16      experimental sites where continuous measurements of wind speed and other environmental
17      conditions are made. This topic is discussed further below in Section 9.6.4.5.
18           No simple exposure index can accurately represent all of the numerous factors operating at
19      different timescales that affect O3 flux into plants and subsequent plant response (Section 9.5).
20      Indices of peaks, such as the 2FIDM are not well suited for discerning exposure-response
21      relationships, because they  do not capture the effects of lower O3 concentrations nor the
22      cumulative effects of O3 on vegetation (Heck and Cowling, 1997; U.S. Environmental Protection
23      Agency, 1996).  For this reason, peak indices have not been used in recent decades to develop
24      exposure-response relationships for vegetation.  Fortunately, other simple indices have shown
25      substantial correlation with responses such as crop yield under experimental conditions. During
26      the 1980s, the most commonly used indices for expressing O3 exposure were 7-, 8-, or 12-h
27      daytime average values over the duration of O3  exposure, which was often 3 months or
28      somewhat less for experimental studies with crops. These indices perform reasonably well for
29      interpreting experimental data on the response of vegetation to ozone, particularly for individual
30      experiments, although they do not explain all of the variation among experiments in
31      retrospective analysis of multiple experiments (Lesser et al.,  1990).

        January 2005                              9-193       DRAFT-DO NOT QUOTE OR CITE

-------
 1           Since the 1980s, cumulative indices such as the SUM06, AOT40, or W126 that
 2      preferentially weight higher concentrations have been used in conjunction with mean indices for
 3      developing exposure-response relationships (Tables 9-15 and 9-16, and Figure 9-1).  Such
 4      indices are often more suitable than mean values because they are cumulative and because they
 5      preferentially weight higher concentrations. Thus, these indices generally provide somewhat
 6      better fits to experimental data than do mean indices, especially in retrospective analyses of
 7      multiple experiments on multiple species (Lee et al., 1994a, 1994b; Lee and Hogsett, 1999;
 8      Tingey et al., 1991).  Unfortunately, no single index has been used consistently even in the
 9      recent literature, making it difficult to compare results among and between experiments and with
10      ambient exposure data. However, Tables 9-14 and 9-17 provide summaries of ambient exposure
11      data for several indices that can be compared to the experimental results reviewed in the
12      remainder of this section. Of the cumulative indices that preferentially weight higher
13      concentrations, the SUM06 index has been used most commonly in the U.S. literature, and it was
14      selected in a meeting of scientific experts on O3 effects  on vegetation as suitable for a secondary
15      standard to protect vegetation (Heck and Cowling, 1997). However, it should be noted that the
16      W126 index has been selected for use in protecting vegetation in Class 1 areas (Federal Land
17      Managers' Air Quality Related Values Workgroup (FLAG), 2000). Even in recent studies, O3
18      data are often presented using only a seasonal mean index value, and so mean values are
19      frequently presented in this section. Such reporting of mean indices should not be interpreted as
20      a preference for them, but rather as a limitation in the data reported in the literature.  Additional
21      information about O3 exposure for individual experiments, including the number and type of O3
22      treatments (addition of a constant concentration of O3 or an amount proportional to ambient
23      levels), and duration, are reported in Tables 9-16 through 9-19.
24           Since the 1996 document, the use of the AOT40 index has become quite common in
25      Europe for identifying and mapping areas of exceedance, but it has not been used much  in the
26      United States. Thus, studies reporting O3 exposure only as AOT40 values are presented in tables
27      summarizing effects on annual, herbaceous perennial, and woody vegetation.  However, such
28      studies are not as commonly cited in the text of this section because AOT40 summary data on O3
29      exposures in the United States are rarely available. This lack makes it difficult to compare
30      experimentally derived exposure-response data expressed as AOT40 to ambient U.S. O3
        January 2005                              9-194       DRAFT-DO NOT QUOTE OR CITE

-------
3
Table 9-16. Summary of Selected Studies of Ozone Effects on Annual Species
to
o
o










VO
1
VO



o
£
H
6
O
2
0
H
O
O
w
0
o
H
W
Species
Bean, cv.
Pros

Bean, cv. Lit

Corn


Cotton, cv.
Deltapine
Cotton, cv.
Deltapine
Oat, cv. Vital

Potato2

Rape, oilseed

Rice, cvs.
Koshi-hikari,
Nippon-bare
Soybean
Soybean, cv.
Essex


Facility
OTC


OTC

OTC


OTC

OTC

OTC

OTC

Open Air

OTC
OTC
OTC



Location
The Netherlands


Germany

Beltsville, MD


Raleigh, NC

Raleigh, NC

Ostad, Sweden

6 sites N. Europe

Northumberland,
UK

Japan
Beltsville, MD
Raleigh, NC



O3 Concentration
(Units are ppb unless
otherwise specified)1
CFtoCF75:
9-hmean = 3-70,
AOT40 = Oto 17.7ppm«h
CF,NF, CF-lx,CF-2x:
mean=l, 14,15,32
CF, +40: 7-h mean = 20, 70


CF, 1.5x:
12-hmean = 21,71
CF,NF, 1.5x,
12-hmean = 24, 51, 78
CF,NF: 7-h mean =12, 27

AOT40 = 6-27 ppm«h

AA, +O3: 7-h mean for
17 days Aug. -Sept = 30, 77,
for 32 days in
May-June = 31, 80
CF, lx, 1.5x,2x,2.75x:7-h
mean= 13.5-93.4
CF, +40: 7-h mean = 25, 72
CF, 1.5 x: 12-h mean for
3 years = 23, 82


Duration
62 days


3 months

1 years


1 years

1 years

1 years

2 years (1 year
at 2 sites)
17 days in fall,
overwinter,
32 days in spring
3 years
2 years
3 years



Variable
g=Green pod
yield

Pod yield

Grain yield


Seed-cotton
weight
Seed-cotton
weight
Grain yield

Tuber yield

Seed yield

Grain yield
Seed yield
Seed yield



Response
(Decrease from
lowest, %)
29 at 9-h mean = 44
(AOT40 = 3.6ppm«h)

56 (CF, 2x)

13


22

21,49(NF, 1.5x)

+2 (n.s.)

4% average for
all experiments
14

3 to 10 at 40 ppb
25
41



Reference
Tonneijck and
Van Dijk (1998)

Brunschon-Harti
etal. (1995)
Mulchi et al.
(1995)Rudorff
etal. (1996c)
Heagle et al.
(1999)
Heagle et al.
(1999)
Pleijel et al.
(1994a)
Craigon et al.
(2002)
Ollerenshaw
etal. (1999)

Kobayashi et al.
(1995)
Mulchi et al.
(1995)
Fiscus et al.
(1997)



-------
3
to
o
o









VO
1
VO
ON



O
^
H
6
o
0
H
O
o
0
o
H
W
Table 9-16 (cont'd). Summary of Selected Studies of Ozone Effects on
Species
Soybean, cvs.
Forrest, Essex
Soybean, cv.
Essex
Soybean, cv.
Essex
Soybean, cv.
Holladay
Soybean, cv.
NK-6955
Soybean,
3 cvs.
Soybean,
3 cvs.
Soybean, cv.
Essex
Soybean, cvs,
Essex, Forrest
Timothy
Watermelon




Facility
OTC

OTC

OTC

OTC

OTC

OTC
OTC

OTC

OTC

OTC
OTC




Location
Maryland

Raleigh, NC

Raleigh, NC

Raleigh, NC

Raleigh, NC

Raleigh, NC
Raleigh, NC

Raleigh, NC

Beltsville, MD

Sweden
Spain




O3 Concentration
(Units are ppb unless
otherwise specified)1
CF, +40: 7-h mean = 24 and
24, 63 and 62 for each year
CF,NF, 1.5x:
12-h mean = 20, 50,79
CF,NF, 1.5x:
12-h mean =18, 42,69
CF,NF, 1.5x:
12-h mean = 18,42,69
CF,NF, 1.5x:
12-h mean = 18,42,69
CF,NF, 1.5x:
12-h mean =14, 36,64
CF,NF, 1.5x: 12-h mean =
24, 49, 83
CF,NF, 1.5x:
12-h mean = 20, 50, 79
CF,NF+: 7-h mean = 24, 58

AOT40 = 10, 20, 340;
12-h mean = 20, 152
CF(O3 = 0),NFml988
AOT40 = 5.96 ppm«h,
SUM06 = 0.29 ppm«h, in
1989 AOT40 =18.92 ppm«h,
SUM06 = 4.95ppm«h



Duration Variable
2 years Seed yield

1 year Seed yield

1 year Seed yield

1 year Seed yield

1 year Seed yield

3 months Seed yield
4 months Seed yield

4 months Seed yield

134 days Seed yield

1 year Biomass
2 expts of 1 year Fruit yield




Annual Species
Response
(Decrease from
lowest, %)
10, 32 (2 cvs.)

16, 37 (NF, 1.5x)

15, 40 (NF, 1.5x)

22, 36 (NF, 1.5x)

+46, +4(NF, 1.5x)

At ambient = +14, 11,
16 for 3 cvs.
At ambient = 17, 13,
18 (3 cvs.)
ll,22(amb., 1.5x)

Essex = +11 (ns),
Forrest = 21
58
19, 3 9 (2 expts)





Reference
Chernikova
et al. (2000)
Heagle et al.
(1998)
Heagle et al.
(1998)
Heagle et al.
(1998)
Heagle et al.
(1998)
Miller et al.
(1994)
Miller et al.
(1994)
Miller et al.
(1998)
Robinson and
Britz (2000)
Danielsson et al.
(1999)
Gimeno et al.
(1999)





-------
Table 9-16 (cont'd).  Summary of Selected Studies of Ozone Effects on Annual Species
to
o
o











1
^5




o
H
6
o
0
H
O
o
0
o
HH
H
W
Species Facility
Wheat1, cv. OTC
Minaret





Wheat1 OTC

Wheat, cv. OTC
promessa

Wheat, cv. OTC
promessa

Wheat, cv. OTC
promessa


Wheat, cvs. OTC
Massey,
Saluda
Wheat, cv. OTC
Turbo
Wheat, cv. OTC
Turbo
Wheat, cv. OTC
Turbo

Location
8 sites in
N Europe





Sweden

SE Ireland


SE Ireland


SE Ireland



Beltsville, MD
Germany
Germany
Germany


O3 Concentration
(Units are ppb unless
otherwise specified)1
12-h mean (SD) low = 26.3
(12.2), 12-h mean (SD)-high
= 51.4 (18.3)
AOT40 mean (SD)
low = 6.18(8.54)ppm«h,
AOT40 mean (SD)
high = 28.23 (23.05) ppm«h
AOT40 0 to 15 ppm«h

CF, +50:
12-h total = 5.6, 32.6 ppm«h

CF, +25:
12-h total = 6.2, 33.4 ppm«h

CF, +25, +50:
12-h total = 6.7, 34,
34 ppm«h

CF, +40:
7-h mean =19, 20 and 61,
65 (2 years)
8-h mean = 5. 9, 61.2, 92.5
8-h mean = 4. 7, 86.4
7-h mean = 5, 41,73


Duration
13 studies of
1 year each





7 years

3 h/day,
5 d/week,
7 weeks
6 h/day,
5 d/week,
7 weeks
+25 = 6 h/day,
5 d/week,
+50 = 3 h/day,
5 d/week,
both 7 weeks
2 years
1 year
1 year
1 year


Variable
Grain yield






Grain yield

Grain yield


Grain yield


Grain yield



Grain yield
Grain yield
Grain yield
Grain yield


Response (Decrease
from lowest, %)
13 (n.s.)






23 at AOT40 =
15 ppm«h
53


+17


Amb + 25 = 3 (n.s.);
Amb + 50 = 17


20
14,40
(mid, high O3)
20
35


Reference
Bender et al.
(1999)Hertstein
etal. (1999)




Danielsson et al.
(2003)
Finnan et al.
(1996a)

Finnan et al.
(1996a)

Finnan et al.
(1996a)


Mulchi et al.
(1995);Rudorff
etal. (1996b)
Bender et al.
(1994)
Bender et al.
(1994)
Fangmeier et al.
(1994)


-------
3
«s
to
o
o











I i-\
M-J
1
VO
oo
o
H
1
O
o
0
H
O
o
0
o
H
W
Table 9-16 (cont'd). Summary of Selected Studies of Ozone Effects on Annual Species
O3 Concentration
(Units are ppb unless Response (Decrease
Species Facility Location otherwise specified)1 Duration Variable from lowest, %)
Wheat, OTC Raleigh, NC 1 2-h mean = 27, 47, 90 2 months Grain yield 5 (n.s.)
winter, 8 cvs.
Wheat, OTC Raleigh, NC 1 2-h mean = 22, 38, 74 2 months Grain yield 16 (n.s.)
winter, 8 cvs
Wheat, cv. OTC Finland 1992: 12-hmean = 14, 30, 2 years Grain yield At highest O3 =
Drabant 61; AOT 40 = 16.3, 34.8, 13 each year
54.6ppm«h. 1993:
12-hmean =9, 21,45;
AOT40 = 10.2, 24.8,
40.6 ppm«h
Wheat, cv. Open Air Northumberland, AOT40 for Mar to 1 year, Grain yield 1 3
Riband UK Aug 93 = 3.5, 6.2 ppm«h overwinter
Values for ambient or NF treatments are indicated in bold.
2
Bold indicates that multiple experiments (more than just 2 years at a single site) were included in the analysis.
















Reference
Heagle et al.
(2000)
Heagle et al.
(2000)
Ojanpera et al.
(1998)




Ollerenshaw and
Lyons (1999)
















-------
3
«s
to
o
o













VO
1
VO
VO



o

rr5
H
M
\M/
o
2!
•£—\
0
H
O
C|
o
1
w
o

o
1 — I
H
W
Table 9-17. Summary of Selected Studies of the


Species
Alfalfa, cvs.
Apica, Team

Bahia grass


Bent grass
(Capillaris sp.)




Blackberry



Clover, white


Clover, white



Clover, white





Clover, white
and red


Clover, white,
cv. Menna



Facility Location
OTC Quebec,
Canada

OTC Auburn, AL


OTC United
Kingdom




Large Alabama
OTC


Ambient MA, OR, NC,
air CA (2 sites)
andVA
Ambient 14 European
air sites


OTC United
Kingdom




OTC Switzerland



OTC Italy


O3 Concentration
(Units are ppb unless
otherwise specified)1
12-hmean: 1991 = 6,39,49,110;
1992 = 0,34,42,94

12-hmean = 22, 45, 91


AOT40 = 0.8-15.0ppm-h





1994: AOT40 = 2-1 12 ppm-h,
SUM06= l-162ppm-h,
1995: AOT40 = 3-83 ppm-h,
SUM06 = 0-132ppm-h
SUM06 for 6-h/day = 10.2-39.4
ppm-h, AOT40 for 12-h/day =
0.6-50.1 ppm-h
AOT40 for 28 d = 0-12 ppm-h



AOT40 = 0.8-15.0ppm-h





CF, NF, NF+. NF++:
12-hmean = 21, 39, 47, 65


CF,NF: AOT40 = 0.1;
8.9 ppm-h, 7-h mean = 24, 53

Effects of Ozone on Perennial Herbaceous Plants


Duration
3 months in
each of
2 years
24 weeks


8 h/day for
3 months




7 months
in 1994,
6 months
in 1995
2 growing
seasons

3 growing
seasons


8 h/day for
3 months




3.5 months/
year for
2 years

2 months




Variable
Biomass


Biomass at ambient O3 for
1 st, 2nd cutting of early
and late season plantings
Biomass, in competition
with 3 other spp. Total
biomass in uncut pots,
aboveground biomass in
cut pots (cut every
14 days).
Percent canopy cover
(grown in old field
community), biomass ripe
fruit number
Biomass ratio
(sensitive/resistant)

Biomass ratio
sensitive/resistant)


Biomass, in competition
3 other spp. Total biomass
in uncut pots, aboveground
biomass in cut pots (cut
every 14 days).

Biomass, in managed
pasture


Biomass



Response (Decrease
from lowest, %)
ForNF: Apica = 31,
21; Team = 14,2
(n.s.)
34,29(n.s.),+6(n.s.),
9 (n.s.)

8 (uncut), +18 (cut)





+124 for cover, n.s.
for biomass, 28% for
ripe fruit number but
sig. chamber effect.
4 at 6-h SUMO6 =
39.4 ppm-h; 12 h
AOT40 = 50.1ppm-h
5atAOT40for
28 days = 0.9-1. 7
ppm-h

18 (uncut), 40 (cut)





24, 26, 52



20




Reference
Renaud et al.
(1997)

Muntifering
et al. ( 2000)

Ashmore and
Ainsworth
(1995)



Barbo et al.
(1998)
Chappelka
(2002)
Heagle and
Stefanski
(2000)
Mills et al.
(2000)


Ashmore and
Ainsworth,
(1995)



Fuhrer et al.
(1994)


Fumagalli
etal. (1997)


-------
3
«s
to
o
o














VO
1
to
o
o




o

rr5
H
o

0
H
iO

O
m
rn
O
O
HH
H
W
lauie ?-i / ^


Species Facility
Fescue, red OTC





Lespediza, OTC
Sericea

Little bluestem OTC


Phleum alpinum OTC



Rasberry OTC


Speedwell, OTC
Germander




Strawberry OTC


Sumac, winged OTC



Timothy OTC


com u>. CMIIIII


Location
United
Kingdom




Auburn, AL


Auburn, AL


Sweden



Ontario


United
Kingdom




United
Kingdon

Alabama



Sweden


nary 01 aeiecieu aiuuies
O3 Concentration
(Units are ppb unless
otherwise specified)1
AOT40 = 0.8-15.0ppm-h





CF,NF,2x: 12-hmean = 23,
40, 83, SUM06 = 0.2, 9.1, 61.0,
AOT40 = 0.6, 7.0, 39.8
CF,NF,2x: 12-hmean = 23,
40, 83 ppb, SUM06 = 0.2,9.1,
61.0, AOT40 = 0.6, 7.0, 39.8
AOT40 = 0.01, 0.02, 0.34 ppm-h;
12-hmean = 20, 152


lx,+12+24: Ambient < 4


AOT40 = 0.8-15.0ppm-h





8-h mean = 27, 92;
AOT40 for +O3 = 24.59 ppm-h

SUM06 = 0 to 132 ppm-h



CF, NF, CF+:
AOT40 = 0.0, 1.3, 20.3 ppm-h;
12-hmean = 20, 68, 152
01 me i^iiecu


Duration
8 h/day for
3 months




10 weeks


10 weeks


1 year



1 day /week
for 7 weeks

8 h/day for
3 months




69 days


6 months



1 year


> 01 uzoiie on reremii


Variable
Biomass, in competition
with 3 other spp. Total
biomass in uncut pots,
aboveground biomass in
cut pots (cut every
14 days).
Biomass


Biomass


biomass



Fruit yield


Biomass, in competition
with 3 other spp. Total
biomass in uncut pots,
aboveground biomass in
cut pots (cut every 14
days).
Fruit size, yield


Percent canopy cover
(grown in old field
community)

Biomass


ai neruaceous riams

Response (Decrease
from lowest, %) Reference
+ 30 (uncut), Ashmore and
+13 (cut) Ainsworth,
(1995)



n.s. Powell et al.
(2003)

n.s. Powell et al.
(2003)

87 (Danielsson
etal. (1999)


No effect at + 12, Sullivan et al.
52 at +24 in only one ( 1 994)
of two cultivars
14 (uncut), 26 (cut) Ashmore and
Ainsworth
(1995)



Size = 14, Drogoudi and
yield = (n.s.) Ashmore
(2000)
95 (Barbo et al.
(1998)


ns in NF, 58 in CF+ Danielsson
etal. (1999)

1 Values for ambient or NF treatments are indicated in bold.
2 Bold indicates that multiple
experiments (more
than just 2 years at a single site)
were included in the analysis.

-------
3
to
o
o



VO
to
o

o
H
6
O
0
H
O
O
0
o
H
W
Table 9-18. Summary of Selected Studies of Ozone Effects on Deciduous Trees and Shrubs
Species Age Facility
Ash, OTC
European
Ash, Seedling OTC
European
Aspen Cutting, OTC
Seedling
Aspen Cutting OTC
Aspen Cutting FACE
Aspen Cutting Large OTC
Aspen First year OTC
Beech, OTC
European
Beech, Seedling OTC
European




O3 Concentration
(Units are ppb unless
Location otherwise specified)1 Duration
Hampshire, UK NF, NF+: Mean = 17.7, 3 years for
44.1; AOT40 for 24h = 1.9, day 100 to
59.9 ppm-h day 162
Switzerland 0.5x, 0.85x, lxs 0.5x+30: 5 months
AOT40 = 0.1,3.4, 7.1,
19.7ppm-h
Michigan CF, lx, 2x; 3 months 7-h 98 days
mean for 1990 = 7-69;
for 199 1=22-92
Michigan CF, lx,2x: 98 days
SUMOO= 11,58, 71 ppm-h
Wisconsin ambient, +90 (exposure 3 years
data not reported)
New York lx, 1.7x, 3x: 92 days
SUM06 = 1, 20, 62 ppm-h;
9-h mean = 40, 74, 124
Pennsylvania 8-h mean = 39, 73 11 weeks
Switzerland 0.5x, 0.85x, lx; 0.5x+30: 5 months
AOT40 = 0.1,3.4, 7.1,
19.7 ppm'h
Belgium CF, NF, +30: 23 April -
8-h mean = 5. 29, 33; 30 Sept
AOT40 = 0. 4.06,
8.88 ppm-h




Variable
Growth and
biomass of
organs
Biomass
Total
biomass
Total
biomass
Volume
(d2*h)
Shoot
biomass
Biomass
Biomass
Growth




Response
(Decrease from
lowest, %)
n.s.
26 at lx, 50 at
O.Sx + 30
nsat lx,29at
2x for all clones
in each year
25-38 at 1 x
21 at +90
14, 25 for
2 clones at 1.7x
14-30 for 3 of
6 N treatments
6 at lx, 30 at
O.Sx+30
No effect




Reference
Broadmeadow
and Jackson
(2000)
Landolt et al.
(2000)
Kamosky et al.
(1996)
Dickson et al.
(2001)
Isebrands et al.
(2001)
Yunand
Laurence
(1999a)
Pell et al.
(1995)
Landolt et al.
(2000)
Border et al.
(2000a)





-------
3
to
o
o












VO
to
o
to




o
>
H
6
O
2
0
H
O
O
0
o
H
W
Table 9-18 (cont'd). Summary of Selected Studies of Ozone Effects on Deciduous Trees
O3 Concentration
(Units are ppb unless
Species Age Facility Location otherwise specified)1
Beech, European Seedling Growth chamber Belgium CF, CF+40, CF+100:
SumO = 0.48, 8.93,25.14
ppm-h; AOT40of
NF+100 = 13.91 ppm-h;
uptake = 159,29657095
molm2
Beech, European 0-3 years OTC Switzerland AOT40 for 24 h/days =
4-73 ppm'h

Beech, Japanese 4 years Growth chamber Japan CF, +60 ppb for 7 h/day

Birch, silver Sapling FACE Finland AOT40 = 1, 15 ppm-h;
7-h mean = 26, 40
Birch, silver Sapling OTC Sweden NF, NF+, NF++, daylight
mean 1997 = 29, 37, 54;
1998 = 25, 42, 71 ppb;
AOT40 1997 = 2.4, 6.9,
35.1; 1998 = 0.6,19.6,
74.7 ppm'h
Birch, Seedling Chamber in Norway AOT40 = 0.1, 2.5, 7.1,
[B. pubescens] glasshouse 7.4, 17.8, 19.8 ppm'h

Cherry, black 2 years OTC Norris, TN CF, lx,2x:
7-h mean = 2 1,50, 97
Cherry, black Seedling OTC GSMNP1 CF, lx, 1.5x, 2x:
SUM06 = 0-40.6 ppm-h,
AOT40 = 0.03-28.3
ppm-h


Duration Variable
7 episodes of Biomass,
5 days diameter




1-3 years Total
biomass

156 days Total
biomass
5 years Biomass

2 years Total
biomass




40 days Biomass

April to August Biomass
76 days Biomass



and Shrubs
Response
(Decrease from
lowest, %)
No effect





20atAOT40
for 24 h = 32
ppm'h
19

34 for root, ns
for stem
Total biomass
nsatNF+, 22 at
NF++; root
biomass 30 at
NF++

Sig. decrease in
rootatAOT40
= 2.5 ppm-h,
shoot at
7.1 ppm'h
No effect
n.s. at lx and
1.5x, 38at2x




Reference
Border et al.
(2001)




Braun and
Fluckiger
(1995)
Yonekura
etal. (2001)
Oksanen
etal. (2001)
Karlsson
et al. (2003)




Mortensen
(1998)

Samuelson
(1994)
Neufeld et al.
(1995)




-------
 VO

 to
 o
 H
 6
 o

 o
 H
O
Species Age Facility
Cherry, black Seedling OTC
Cherry, black 1 year OTC
Cherry, black Seedling OTC
Cottonwood, Cutting Ambient,
Eastern in buried pots
with irrigation
Grape 3 years OTC
Oak Seedling OTC

Oak, red Seedling OTC

O3 Concentration
(Units are ppb unless
Location otherwise specified)1
GSMNP1 CF, 0.5x, Ixj.Sx^x:
SUM06 = 0-53.7 ppm-h;
AOT40 = 0-40.4 ppm-h
Delaware, OH CF, 0.5, 1, 1.5, 2x:
SUMOO in 1990 = 17-107
ppm-h, in 1991 = 31-197
ppm'h
Pennsylvania CF, 0.75x, 0.97x:
7-h mean = 39 to 46,
SUM06 = 0-10.34 ppm-h
In and within 100 12 h mean = 23-49 ppb
km of New York
City, NY
Austria CF, lx, +30, +50:
(AOT40 = O-50 ppm-h
Hampshire, UK NF, NF+:
Mean = 17.7, 44,
AOT40for24h=1.9,
59.9 ppm-h
Norris, TN SUM06 for 3 years = 0,
29, 326 ppm-h; SUMOO
for 3 years = 147, 255 and
507 ppm'h
Duration
140 days
2 years (in 1990
for 3.5 months,
1991 for
4 months)
3 years for
17 weeks
2 months
each year,
3 10-years
experiments
2 years
(preflowering,
past harvest)
3 years for day
100 to day 162

3 years

Variable
Biomass
Total
biomass
Total
biomass
Total
biomass
Fruit yield
Biomass of
organs

Total
biomass

Response
(Decrease from
lowest, %)
n.s. at lx and
1.5x, 59at2x
no effect at 1 x
and 1.5x, 32 at
2x
6 at 0.75 x, 14 at
0.97x
33% decrease at
38 ppb
compared to
23 ppb
Calculated 10
atAOT40 =
27 ppm'h
30 for total
biomass

n.s.

Reference
Neufeld et al.
(1995)
Rebbeck
(1996)
Kouterick
et al. (2000)
Gregg et al.
(2003)
Soja et al.
(1997)
Broadmeado
wand
Jackson
(2000)
Samuelson
etal. (1996)

o
HH
H
W

-------
fa
g Table 9-18 (cont'd). Summary of Selected Studies of Ozone Effects on Deciduous Trees
to
o
O Species Age Facility
Oak, red 30 years OTC



Maple, red 2 years OTC

Maple, sugar 1 year OTC


Maple, sugar Seedling Large OTC

VO
to
o
Maple, sugar Seedling Large OTC




o
5d Plum, Casselman Sapling Large OTC
M
H
§ Poplar, black Seedling OTC
*
0
o
O3 Concentration
(Units are ppb unless
Location otherwise specified)1
Norris, TN SUM06 for 3 years = 0,
29, 326 ppm-h; SUMOO
for 3 years = 147, 255 and
507 ppm-h
Norris, TN CF, lx,2x:
7-h mean = 2 1,50,
97 ppm-h
Delaware, OH CF, 0.5, 1.5, 2x:
UMOOinl990=17to
107ppm-h, in 1991 =31
to 1 97 ppm'h
Ithaca,NY CF, lx, 1.5x,2x: 3 years
SUMOO = 148 to
591 ppm-h; daytime
mean= 19. 7 to 40. 7
Ithaca,NY lx,1.7x, 3x:
3 years 12-h mean = 38,
69,117


Fresno, CA CF, lx,+O3:
12-h mean =31, 48, 91

Belgium CF, NF, +30:
8-h mean = 5, 29, 33;
AOT40 = 0, 4,8.9 ppm-h

Duration
3 years



April to August

2 years (in 1990
for 3.5 months,
1991 for
4 months)
3 years for 134,
128, 109 days

3 years for 109,
143, 116 days



4 years


23 April -
30 Sept

Variable
Stem
increment


Biomass

Total
biomass


Biomass


Total
biomass



Stem
increment,
fruit yield
Diameter,
height

and Shrubs
Response
(Decrease from
lowest, %)
n.s. despite 50%
reduction in net
photosynthesis

No effect

n.s., but linear
trend


No effect


For 1.7x and
3x: 21, 64 in
low light, 26
and 41 in high
light
Fruit yield 16 at
lx, stem +14 at
+03
29 for diameter
in NF+, no
effect on height


Reference
Samuelson
etal. (1996)


Samuelson
(1994)
Rebbeck
(1996)


Laurence
etal. (1996)

Topa et al.
(2001)



Retzlaffetal.
(1997)

Bortier et al.
(2000b)

O
HH
H
W

-------
fa
3
fa
to
O
(^


Table 9-18 (cont'd).
Species Age Facility
Poplar, hybrid 0 year FACE
(P. tremuloides
x P. tremula)
Poplar, hybrid Cutting OTC
Yellow-poplar 1 year OTC
Summary of Selected Studies of Ozone Effects on Deciduous Trees and Shrubs
O3 Concentration
(Units are ppb unless
Location otherwise specified)1
Finland AOT40 = 0.07, 1.6
ppm'h; 7-h mean = 30, 38
Michigan CF, CF+100:
12, 48 ppm-h
Delaware, OH CF, 0.5, 1.5, 2x:
SUMOOinl990 = 17to
107ppm-h, in 1991 =31
to 1 97 ppm-h
Duration
2 months
60 days
2 years (in 1990
for 3.5 months,
1991 for
4 months)
Variable
Biomass,
height
Total
biomass
Total
biomass
Response
(Decrease from
lowest, %)
n.s. for biomass,
6 for height
46 for average
of 5 clones
No effect
Reference
Oksanen
etal. (2001)
Dickson et al.
(1998)
Rebbeck
(1996)
 VO

 to
 o
 H
 6
 o

 o
 H
O
          1 Values for ambient or NF treatments are indicated in bold.

          2 Bold indicates that multiple experiments (more than just 2 years at a single site) were included in the analysis.
o
HH
H
W

-------
3
3
to
o
^ Species
Fir,
Douglas
Hemlock,
eastern
Pine,
loblolly
Pine,
loblolly
Table 9-19. Summary of Selected Studies of Ozone Effects on Evergreen Trees and Shrubs
Age Facility Location
Seedling Open air British
Columbia
Seedling OTC GSMNP1,
TN
12 weeks OTC Oak Ridge,
TN
1 year OTC Alabama
O3 Concentration
(Units are ppb unless
otherwise specified)1
12 tits:
12-h mean 1988 =18-41;
1989 = 27-66
CFto2x:
SUM06 = 0.2-108.1 ppm«h,
AOT40 = 0.2-63. 9 ppm«h
CFto2x:
24-h summer = 74, 137, 169,
206, 284 ppm«h
1994:
AOT40 = 2-112ppm«h,
Duration
1988 = 92
days; 1989 =
101 days
3 years
3 months
2 years, April
to October
Variable
Second flush
biomass
Biomass
Biomass
Dry weight,
height,
Response
(decrease from
lowest, %)
Calculated 55 at
highest exposure
No effect
14 in lx (avgfor all
families)
n.s.
Reference
Runeckles and
Wright (1996)
Neufeld et al.
(2000)
McLaughlin
etal. (1994)
Barbo et al.
(2002)
VO
SUM06
1995:
                                                                                                diameter
                                                                       «h,
o
Oi




O

£j
H
6
o
2;
0
H
O
o


o
H
W

Pine, 3 years
loblolly



Pine, Seedling
ponderosa



Pine, 39 to
ponderosa 45 years

Pine, Seedling
ponderosa




SUM06 = 0-132ppm«h
OTC Raleigh, NC Ambient, CF, NF, 1 .5 x .,
2.5x: 12-h mean = 54, 29, 47,
76,98


OTC Corvalhs, For CF 12-h SUM06 =
OR 0 ppm«h; for +03 12-h
SUMO6 = 22, 27, 31ppm«h
for 3 years

Ambient CA 24-h mean for 3 weeks late
gradient July and early August for
1993 and 1994 = 70-90 ppb
OTC CA CF, lx,2x:
24-h mean
approx. 20, 60, 120




5 months




3 years:
16 weeks,
16 weeks,
14 weeks

Ambient
gradient








Height,
diameter,
needle length


Total biomass




Fine and
medium root
growth
Total biomass






No effect on stem
height or diameter,
decrease in needle
length

No effect without
grass, 25 with grass
present


85 at most polluted
site.

n.s.






Anttonen et
(1996)



Andersen et
(2001)



Grulke et al.
(1998)


al.




al.







Takemoto et al.
(1997)










-------
3
«s
to
o
o











VO
1
o



M
\^
>
H
6
o

0
H
O


0
o
H
W
Table 9-19 (cont'd). Summary of Selected Studies of Ozone Effects on Evergreen Trees and Shrubs


Species
Pine,
Scots

Pine,
Scots
Pine,
Scots

Pine,
Scots

Pine,
Scots

Pine,
Table
Mountain
Pine,
Virginia




Sequoia,
giant


Spruce,
Norway




O3 Concentration
(Units are ppb unless
Age Facility Location otherwise specified)1
OTC Hampshire, NF, NF+:
UK Mean = 17.7, 44.1; 24-h
AOT40 = 1.9, 59.9 ppm«h
3 -6 years Free air Finland Amb, +O3:
AOT40 = 0-1, 2-13 ppm«h
Seedling OTC Switerzland 0.5x, 0.85x, lx, 0.5x+30:
AOT40 = 0.1, 3.4,7.1,
19.7 ppm«h
3 years OTC Finland CF, lx,+O3:
24 h AOT40 for
2 years = 0.5, 6, 73 ppm«h
3 years Free air Finland lx,+O3:
24 h AOT40 for 2 years = 2,
37 ppm«h
Seedling OTC GSMNP2, CFto2x:
TN SUM06 = 0.2-116.4 ppm«h,
AOT40 = 0.2-71. 7 ppm«h
Seedling OTC GSMNP2, CFto2x:
TN SUM06 = 0.1-32.8, 47.9,
56.2 ppm«h;
AOT40 = 0.1-19.3, 27.1, 34.4
ppm«h

125 years Branch California 0.25x, lx; 2x, 3x;
chamber 24-h SUMOO approx. 10, 85,
180, 560ppm«h

4-7 years Open air Finland Amb, +O3:
AOT40 = 01,2-13ppm«h






Duration
3 years for
62 days

3 years

5 months


2 years
(4 months
each)
3 years
(3-4 months
each)
3 years


1 -2 years
(3 expts)




61 days



3 years





Response
(decrease from
Variable lowest, %)
Total biomass 15


Biomass No effect

Biomass 14atlx, 22 at
0.5x+30

Biomass No effect


Root and shoot 32 only for root
biomass biomass in high N
treatment
Biomass Slight decrease in
older needle mass
only
Biomass No effect





Branch growth No effect



Biomass No effect







Reference
Broadmeadow
and Jackson
(2000)
Kainulainen
et al. (2000)
Landolt et al.
(2000)

Utriainen et al.
(2000)

Utriainen and
Holopainen
(2001)
(Neufeld et al.
(2000)

Neufeld et al.
(2000)




Grulke et al.
(1996)


Kainulainen
et al. (2000)





-------
                  Table 9-19 (cont'd).  Summary of Selected Studies of Ozone Effects on Evergreen Trees and Shrubs
%
to
o
s












VO

g



a
H
1
O
o
0
H
O


Species Age Facility Location
Spruce, 3 -7 years OTC Sweden
Norway

Spruce, Seedling OTC Switerzland
Norway

Spruce, 0-3 years OTC Switzerland
Norway

Spruce, Seedling OTC Sweden
Norway


Spruce, Sapling Large Ithaca, NY
red OTC


1 Values for ambient or NF treatments are indicated in
2 Great Smoky Mountains National Park.







O3 Concentration
(Units are ppb unless
otherwise specified)1 Duration
CF, 1.5x; 4 years
12-h mean for 4 years = 12,
44; AOT40 = 2, 23 ppm«h
0.5x,0.85x, lx,0.5x+30: 5 months
AOT40 = 0.1, 3.4,7.1,
19.7 ppm«h
AOT40 for 24 h for 1 to 1-3 years
3 years = 22 to 63 ppm«h

CF, lx, 1.5x; 4 years
AOT40 daylight for 4 years =
1,16, 79ppm«h

CF, lx, 1.5x, 2x: 4 years:
total for 4 years = 21 1 to 98-124
569 ppm«h; days/year
daytime mean = 21-71
bold.








Response
(decrease from
Variable lowest, %) Reference
Total biomass 8 Karlsson et al.
(2002)

Biomass n.s. Landolt et al.
(2000)

Total biomass n.s. Braun and
Fluckiger
(1995)
Stem volume No effect in final Wallin et al.
year (2002)


Biomass No effect Laurence et al.
(1997)











O
HH
H
W

-------
 1      exposures. The development of critical levels in Europe has been based primarily on the AOT40
 2      index, so this index is discussed in that context.
 3            In addition to peak-weighting, there is also evidence that the timing of exposure during
 4      plant growth is important. For example, the greatest effects on grain yield are due to exposure
 5      during grain filling, rather than earlier or later in the growing season (Lee et al. 1988; Pleijel
 6      et al., 1998; Soja et al., 2000; U.S. Environmental Protection Agency, 1996; Younglove et al.,
 7      1994). The importance of respite times was also discussed in the previous criteria documents
 8      (U.S. EPA, 1978, 1988, 1996) but remains difficult to quantify (Section 9.5). Even when some
 9      of these aspects of O3 exposure can be elucidated, it is difficult to apply this knowledge to
10      developing exposure-response relationships based on data in the scientific literature, because O3
11      exposure is often reported only in the form of a summary index such as a 12- or 24-h mean,
12      SUM06, or AOT40.
13            Table 9-14 presents summaries of ambient O3 exposure patterns in the United States for
14      1982 to 1981 for several indices including the 7-h mean and SUM06. More recent summaries
15      for the entire United States for these indices are not available,  but Table 9-15 summarizes more
16      recent data for the central and eastern United States. As shown in Table 9-15, from 1989 to
17      1995, mean 12-h 3-month SUM06 values (in ppm-h) at 41 rural sites in the Clean Air Status and
18      Trends Network were 31.5 for the Midwest, 18.9 for the Upper Midwest, 33.2 for the Northeast,
19      13.2 for the Upper Northeast (NH, ME), 34.5 for the South-Central, and 19.2 for the Southern
20      Peripheral subregions (Baumgardner and Edgerton,  1998). These results are important because
21      these sites were selected to represent rural areas, while many other monitoring sites represent
22      urban or suburban areas.  For these same subregions, W126 values ranged from 12.8 to
23      25.6 ppm-h. From 1989 to 1995, O3 concentrations decreased about 5% for daily and 7% for
24      weekly values for most of these sites, after adjusting for meteorological conditions (Holland
25      et al., 1999).  These trends were statistically significant at about 50% of the sites  (p < 0.05).
26      However, because the trend analysis was intended to examine the efficacy of O3 emissions
27      controls, the trends were adjusted for meteorological conditions. Thus, they do not reflect the
28      actual trends in O3 exposure over time.
29
        January 2005                              9-209       DRAFT-DO NOT QUOTE OR CITE

-------
3
Species
o
w Bean, cv. Lit
Table 9-20. Ethylene Diurea Effects on Vegetation Responses to Ozone
Description
10-cm pots in OTCs in
Germany
EDU application
Soil drench 200 mL of
150 ppm solution per
plant every 14 days
Ozone exposure Effects of EDU
CF, NF, CF-1 x, CF-2x : O3 reduced pod, shoot, and
mean= 1, 14, 15, 32 ppb root mass. EDU increased
root, leaf, and shoot mass,
but a significant interaction
with O3 occurred only for
root weight.
Reference
Brunschon-Harti
etal. (1995)
          Bean, cv. BBL-290
2 expts in 5.5 L pots
in OTCs with 4
O3 treatments
                         Soil drench every 14
                         days, in expt 1=0.14,
                         28, 56, 120 mg/L potting
                         medium; expt 2 = 0, 8,
                         16, 32 mg/L
                         2 expts with CF, NF and
                         2 constant additions of
                         O3. 7-h mean O3 (ppb)
                         for Expt 1 = 34, 70, 95,
                         121; Expt 2 = 19, 42.
                         74, 106
Visible injury and reduced
total biomass or yield, even in
CF treatment. Within an O3
treatment, sometimes
increased yield (Expt 2 only).
Miller etal. (1994)
VO
to
          Bean
Pots with potting mix at
3 locations in Spain
(2 years at 1  site)
                         Soil drench of 200 mL of
                         increasing concentrations
                         of 100, 150,200,250
                         ppm every 14 days
                         (4- 10 mg 1:1 soil)
                         AOT40 = 0.4-1.8ppm-h
0 to 50% increase in
pod mass, but did not restore
yield at sites with higher O3.
Ribas and Penuelas
(2000)
          Bean, cv. Lit
Pots with potting mix at
4 sites in the Netherlands
                         Soil drench of 200 mL of
                         increasing concentrations
                         of 100, 150,200,250
                         ppm every 14 days
                         (4- 10 mg 1:1 soil)
                         AOT40 = 0.64-0.98,
                         7-h mean = 49-55 ppb
Average 20% yield increase
at all sites.
Tonneijck and
Van Dijk( 1997)
H
6
o
o
H
O
O
HH
H
W
          Bean, cv. Lit
Pots with potting mix at
1 site in Belgium
Soil drench of 200 mL of   AOT40 = 0.81 ppm-h
increasing concentrations
of 100, 150,200,
250 ppm every 14 days
(4-10 mg 1:1 soil)
                                                                            16% yield increase.
                              Vandermeiren et al.
                              (1995)
Clover, subterranean
Plants in 10-cm pots at
4 rural sites in the
Netherlands for 3 years
100 ml of 150 ppm
solution as soil drench
every 14 days for
2 months
AOT40 = 0-0. 56 ppm-h
for 4-week periods
Injury, but not leaf biomass
was affected by EDU and
O3 exposure.
Tonneijck and
Van Dyk (2002)

-------
to
o
o
                                   Table 9-20 (cont'd).  Ethylene Diurea Effects on Vegetation Responses to Ozone
Species
Clover, white
Clover, white, cv. Menna
Description
15-cm pots in field, well
watered, 12 locations
throughout Europe,
3 years
2 expts,10-cmpots in
field in Italy, see also
companion OTC expt
EDU application
100 ml of 150 ppm
solution as soil drench
every 14 days for
3 months
100 ml of 150 ppm
solution as soil drench
every 14 days for
Ozone exposure
AOT40 (28 days) = 0-
20 ppm-h
AOT40 = 15.5,
12.1 ppm«h;
7-h mean = 69, 60
Effects of EDU
Change in biomass ratio, weak
linear relationship (r2 = 0. 16)
stronger relationship using
ANN and climatic factors
n.s.
Reference
Ball etal. (1998)
Fumagalli et al.
(1997)
                                                         2 months
vo
to
         Clover, white, cv. Menna
         Poplar, hybrid
2 expts,10-cm pots in
OTCs in Denmark
Stem injections, field,
cuttings, 1 or 2 years
Soil drench 100 mL of
150 ppm solution every
14 days

Approx. 125 or 250
mg/leaf (low, high
EDU treatments)
5 times every 14 days
CF, NF, NF+ 25,
NF+50 ppb,
O3 exposure not reported

1991:
7-h mean = 56,
AOT40 = 23;
1992:
7-h mean = 59,
AOT40 = 27
No effect of EDU despite
highly significant effect of
O3 on above-ground biomass

No effect on biomass; 6%,
12% more severely O3
damaged leaves in high
EDU for 2 years
(Mortensen and
Bastrup-Birk, 1996)
Ainsworth et al.
(1996)
         Pine, loblolly
H
6
o
o
H
O
         Radish, cv. Cherry Belle
1 year old half-sib
seedlings in field in TX
for 3 years
Plants in pots in potting
mix exposed for 5 weeks
in southern Sweden.
150, 300, 450 ppm
every 14 days
Soil drench containing
20 mg EDU applied
2 times, 14 days apart
1995, 1996, 1997,
no. h>40ppb= 1723,
2297, 2052;
no. h > 60 ppb = 378,
584, 528;
peak =113, 102, 118
24-h mean = 31 ppb,
7-h mean = 36 ppb,
AOT40 = 1.3 ppm-h.
For EDU 450 trt, above-
ground biomass increased
approx 46% (n.s. in other
treatments
24% increase in hypocotyl
mass,  18% increase in
shoot mass
Manning et al.
(2003)
Pleyeletal. (1999b)
o
HH
H
W

-------
 1      9.6.4  Effects of Ozone on Annual and Biennial Species
 2           Much of the research on short-lived species during the last decade has been conducted in
 3      Europe.  Several European studies have focused on wheat with an emphasis on developing
 4      critical levels, as discussed below in Section 9.6.4.6.
 5           An extensive search of the literature was performed using several electronic databases to
 6      identify scientific articles containing quantitative information on both the amount of O3 exposure
 7      and its effects on vegetation. Greater emphasis is placed on studies with longer duration with O3
 8      exposure concentrations and environmental conditions that were as similar as possible to
 9      ambient conditions. Many of the studies reviewed herein were conducted in OTCs. In the
10      United States, nearly all of such studies have used the type of OTC developed by Heagle et al.
11      (1973). For the few studies in the U.S. that used other types of OTCs, they are described briefly
12      in the text. In Europe, a wide variety of styles of OTCs have been used. See Section 9.2 for
13      further information about the use of OTCs.  The emphasis in this subsection is on quantifying
14      exposure-response relationships for annual plants, with a focus on the response of above-ground
15      biomass and yield of species grown as crops or occurring as native or naturalized species in the
16      United States. Emphasis is placed on studies not included in the 1996 AQCD (U.S.
17      Environmental Protection Agency, 1996), including a few studies published prior to 1996.
18      However, an attempt is made to compare the results of these recent studies of individual species
19      to those reviewed in the 1996 AQCD.
20
21      9.6.4.1 Effects on Growth, Biomass,  and Yield of Individual Species
22           Most research on the effects of O3 on herbaceous species has evaluated growth, biomass, or
23      yield of commercial portions of crop or forage species.  It is well established that reproductive
24      organs such as seeds may be particularly sensitive to injury or biomass reductions due to O3, as
25      reviewed recently by (Black et al. 2000}. As discussed in Section 9.4, numerous analyses of
26      experiments conducted in OTCs and with naturally occurring gradients demonstrate that the
27      effects of O3 exposure vary depending on the growth  stage of the plant. Plants grown for seed or
28      grain are often most sensitive to exposure during the seed or grain-filling period (Lee et al.,
29      1988; Pleijel et al., 1998; Soja et al., 2000; Younglove et al., 1994), whereas plants grown for
30      biomass production, such as alfalfa, may be  sensitive throughout the growth period  (Younglove
31      et al., 1994).  However, because different species are  sensitive during different periods of their

        January 2005                             9-212       DRAFT-DO NOT QUOTE OR CITE

-------
 1      growth and, because planting or germination dates vary throughout large regions even for a
 2      single species, no single phenological weighting scheme can appropriately and practically
 3      represent all vegetation in all locations in the United States. For natural populations, reductions
 4      in seed yield might be particularly important if subsequent seedling establishment is
 5      compromised by O3.
 6           Green beans (cv. Pros) were grown in pots in OTCs in the Netherlands for 62 days and
 7      exposed to 6 treatments consisting of constant O3 additions to charcoal filtered (CF) chambers
 8      (see Section 9.2) for 9 h/day (Tonneijck and Van Dijk, 1998). Bean yield response to O3 was
 9      nonlinear, with an apparent threshold near the CF30 (charcoal filtered with a constant addition of
10      30 ppb O3) treatment with a 9-h mean O3 concentration of 28 ppb and an AOT40 value of
11      0.1 ppm-h). Yield was reduced by 29%  at a 9-h mean value of 44 ppb corresponding to an
12      AOT40 value of 3.6 ppm-h (Table 9-17). Beans were grown in pots in OTCs for 3 months with
13      the following O3 treatments: CF, nonfiltered (NF), CF with O3 added up to ambient, and CF
14      with 2x-ambient O3 (Brunschon-Harti et al., 1995). Ozone reduced pod mass by 56% with a
15      mean concentration of 32 ppb in the 2* ambient treatment as compared with 1 ppb in the CF
16      treatment (daily averaging time not reported). A second treatment factor in this experiment was
17      addition of EDU, as discussed below under the heading "Studies Using Ethylene Diurea as a
18      Protectant". These yield reductions are greater than those previously reported in four similar
19      studies summarized in the 1996 AQCD (Table 5-25 of U.S. Environmental Protection Agency,
20      1996). Greater sensitivity in the more recent experiments may be due to cultivar differences or
21      other differences in experimental protocols.
22           In a study with OTCs  on silty loam soil in Beltsville, MD, corn yield was reduced by 13%
23      with exposure to a 7-h mean concentration of 70 ppb O3 compared to a CF treatment with a 7-h
24      mean concentration of 20 ppb (Mulchi et al., 1995; Rudorff et al., 1996a; Rudorff et al., 1996c).
25      In this study, different amounts of O3 were added above ambient levels for 5 days follows: 20,
26      30, 40, 50,  60 ppb, except that O3 was not added to exceed a total concentration of 120 ppb
27      (Rudorff et al.,  1996c, 1996a).
28           In two studies conducted in Raleigh, NC, cotton (cv. Deltapine 51) was grown in pots and
29      exposed to CF and 1.5*  (nonfiltered, see Section 9.2) O3 in one year, and CF, NF, and
30      1.5 x-ambient O3 in the second year, with ambient and elevated CO2 concentrations (Heagle
31      et al., 1999; Table 9-17). In the first year, yield decreased by 22% with 1.5x-ambient O3 (12-h

        January 2005                             9-213        DRAFT-DO NOT QUOTE OR CITE

-------
 1      mean value of 71 ppb). In the second year, yield decreased by 21% and 49% with exposure to
 2      ambient or 1.5x ambient O3 (12-h mean values of 51 and 78 ppb, Table 9-17). Increased CO2
 3      levels prevented or reduced this yield suppression (Heagle et al., 1999). These yield reductions
 4      are similar to those reported previously in four similar studies summarized in the 1996 AQCD
 5      (Table 5-25 of U.S. Environmental Protection Agency, 1996).
 6           In a study of oats in OTCs in southern Sweden, exposure to ambient (NF) O3 did not affect
 7      grain yield (Pleijel et al., 1994a).  Ambient O3 concentration expressed as a 7-h mean was
 8      27 ppb, with only 1 h greater than 80 ppb and none above 90 ppb.
 9           The interactive effects of elevated O3 and CO2 additions on potato yield (cv Bintje) were
10      studied in OTCs at 6 sites in northern Europe as part of the CHIP (Changing Climate and
11      Potential Impacts on Potato Yield and Quality) program (Craigon et al., 2002). Ozone was
12      added to a target daily average value of 60 ppb, and AOT40 values across all years and
13      experiments ranged from ~6 to 27 ppm-h.  The O3 treatment reduced total tuber yield an average
14      of 4.8% with elevated O3 treatment across all experiments (Craigon et al., 2002).  This total
15      effect was statistically significant even though the effects of individual experiments generally
16      were not (Craigon et al., 2002), due to the increased power of the pooled analysis. Several
17      publications report other aspects of the CHIP experiments or present results of individual
18      experiments (De Temmerman et al., 2002a, 2002b;  Donnelly et al., 2001a, 2001b; Fangmeier
19      et al., 2002; Finnan et al., 2002; Hacour et al., 2002; Lawson et al., 2002; Pleijel et al., 2002;
20      Vandermeiren et al., 2002; Vorne et al., 2002).
21           The effect of an intermittent constant addition of O3 using a free air exposure system in
22      Northumberland, UK was investigated with the oilseed rape cultivar "Eurol" (Ollerenshaw et al.,
23      1999).  Ozone was added for 6 h/day for 17 days. The ambient treatment had a mean value of
24      30 ppb  and the O3 addition treatment had a mean of 77 ppb. After overwintering, O3 was added
25      for 32 days for 7 h/day between May and June (mean values of 31 and 80 ppb). Yield was
26      reduced by 14% despite the lack of any foliar symptoms.
27           Field fumigation chambers ventilated with fans on both ends were used to assess effects of
28      five O3  treatments on rice over 3 years in Japan (Kobayashi et al., 1994; 1995). All  O3
29      treatments used CF air, and O3 was added to the 0.5, 1.0, 1.5, 2.0, or 2.75* ambient
30      concentration for 7 h/day.  Based on a linear regression for the 3 years, yield decreased by 3 to
31      10% at  a 7-h mean concentration of 40 ppb (Table 9-17). This decrease is greater than that

        January 2005                             9-214       DRAFT-DO NOT QUOTE OR CITE

-------
 1     found for rice in earlier studies in California (Kats et al., 1985), although whether this difference
 2     is due to differences in cultivars, experimental treatment, or environmental factors cannot be
 3     determined.
 4           During 3 years in Beltsville, MD, the soybean cultivars Essex and Forrest were exposed to
 5     CF air and NF air in OTCs (Chernikova et al., 2000; Robinson and Britz, 2000) with O3 added as
 6     described for experiments with corn and wheat at Beltsville (Mulchi et al., 1995; Rudorff et al.,
 7     1996c). During 1994 and 1995, as previously found for these cultivars, Essex was less sensitive
 8     than Forrest, with yield decreases of 10% (n.s.: p > 0.1) compared to 32% for Forrest (p < 0.01)
 9     (Chernikova et al., 2000). There was no evidence of water stress in this experiment. In 1997,
10     the two O3 treatments were CF (7-h mean = 24 ppb) and NF with a constant addition of O3 (7-h
11     mean = 58 ppb) (Robinson and Britz, 2000). The yield of Essex was not significantly affected,
12     while the yield of Forrest was decreased by  21% (Table 9-17).
13           In a study in Raleigh, NC the soybean cultivar Essex was grown in pots and exposed to CF
14     and 1.5* ambient O3 concentrations during three growing seasons (Fiscus et al., 1997).  Over the
15     3 years, exposure to an average 12-h mean O3 concentration of 82 ppb reduced soybean yield by
16     41% (Table 9-17). In  similar studies also in Raleigh, NC, Essex was exposed to CF, NF, and
17     1.5x ambient O3 for two seasons (Heagle et al., 1998).  Yield decreased by 16% and 15% in the
18     2 years by ambient O3 (12-h mean values of 50 and 42 ppb), and decreased by 37 and 40% with
19     exposure to 1.5x ambient O3 (12-h mean values of 79 and 69 ppb, Table 9-17). In this same
20     experiment in the second year, similar yield reductions were observed for the cultivar Holladay,
21     while the growth of cultivar NK-6955 was increased substantially by ambient O3 exposure.
22     All three cultivars were grown in the same chambers in this experiment, and the authors
23     suggested that NK-6955 plants may have shaded the other cultivars to some extent.
24           In a 2-year study using OTCs in Raleigh, NC,  the soybean cultivars Coker 6955, Essex,
25     and S53-34' were exposed CF, NF, and 1.5x ambient O3 treatments (Miller et al.,  1994).
26     Seasonal mean 12-h O3 concentrations ranged from  14 to 83 ppb. As compared to the CF
27     treatment, ambient O3  exposure (NF treatment) reduced seed yield by 11 to 18% except for
28     Coker 6955 in the first year (1989) which showed a yield increase of 14%. The 1.5x ambient O3
29     treatment reduced yield by 32 to 56% in all  cultivars in both years.  In a similar subsequent
30     experiment with the cultivar Essex, exposure to a 12-h mean ambient O3 concentration of 50 ppb
       January 2005                             9-215       DRAFT-DO NOT QUOTE OR CITE

-------
 1     reduced yield by 11%, while exposure to 79 ppb reduced yield by 22% (Table 9-17) (Miller
 2     etal., 1998).
 3          These yield reductions for soybean are generally similar to those reported previously in
 4     13 similar studies summarized in the 1996 AQCD (Table 5-23 of U.S. Environmental Protection
 5     Agency, 1996).
 6          A reanalysis of 7 years of data from OTC experiments with wheat in Ostad, Sweden
 7     showed that relative yield linearly decreased with increasing O3, with a maximum yield loss of
 8     23% at an AOT40 value of 15 ppm-h (Danielsson et al., 2003).  A very similar response was
 9     found using the flux (stomatal conductance) model of Emberson et al.  (2000b), and a similar
10     amount of the variance was explained by the flux model (for AOT40 model, r2 = 0.34 and for the
11     Emberson flux model, r2 = 0.39). A modified flux model developed and calibrated for this site
12     also had a similar linear response equation, but explained much more of the variance (r2 = 0.90).
13          During the 1990s, a major European research program investigated the combined effects of
14     CO2, O3, and other physiological stresses on wheat (Bender et al., 1999; Hertstein et al., 1996,
15     1999; Jager et al., 1999). The European Stress Physiology and Climate Experiment ("ESPACE-
16     wheat") program included 13 experiments in OTCs at eight sites in northern Europe over 3
17     years.  Low-and high-O3 exposures in these experiments had the following values:  12-hmean
18     (SD) low = 26.3 ppb (12.2), 12-h mean (SD) high = 51.37 (18.3) ppb, AOT40 mean (SD) low =
19     6.2 (8,5) ppm-h, AOT40 mean (SD) high = 28.3 (23.0) ppm-h, as calculated from data presented
20     in Table 3 of Hertstein et al. (1999).  An analysis of all  13 experiments showed that high O3 at
21     ambient CO2 reduced yield by 13% on average (Bender et al., 1999). However, this reduction
22     was not statistically significant based on an ANOVA, and the authors concluded that the wheat
23     cultivar Minaret may be relatively tolerant to O3 (Bender et al.,  1999). Results of some
24     individual studies within this program have been reported previously (Donnelly et al., 1999;
25     Fangmeier et al., 1996, 1999; Mulholland et al., 1997, 1998b, 1998a; Pleijel et al., 2000b).
26          In a study with OTCs on silty loam soil in Beltsville, MD, wheat yield was reduced by
27     20% on average over 2 years with 7-h mean concentrations of 61 and 65 ppb O3 compared with
28     CF treatment with a 7-h mean concentration of 20 ppb (Mulchi et al., 1995; 1996a; Rudorff
29     et al., 1996c). In the above study, different amounts of O3 were added above ambient (levels for
30     5 days) as follows: 20, 30, 40, 50, 60 ppb, except that O3 was not added to exceed a total
31     concentration of 120 ppb (Rudorff et al., 1996c). Wheat grown in pots in OTCs was  exposed to

       January 2005                            9-216       DRAFT-DO NOT QUOTE OR CITE

-------
 1      elevated O3 and water stress in Germany, and yield was decreased by 35% in the 2x -ambient
 2      treatment with a 7-h mean O3 concentration of 71 ppb statistically significant effects were not
 3      seen in the 1 x-ambient treatment (Fangmeier et al., 1994). In two studies conducted in Raleigh,
 4      NC, soft red winter wheat was grown in pots and exposed to CF, NF, and 1.5x-ambient O3, with
 5      ambient and elevated CO2 concentrations (Table 9-17) (Heagle et al., 2000). In the first
 6      experiment, eight cultivars were exposed to 12-h mean O3 concentrations of 27, 47, and 90 ppb,
 7      and in the second experiment two of these cultivars were exposed to 22, 38, and 74 ppb.  There
 8      was a trend toward decreased yield in both experiments, but these trends were not statistically
 9      significant.  The wheat cultivar Drabant was exposed to CF, NF, and a constant addition of
10      35 ppb during 1992 and 1993 using the Heagle-type OTCs (Heagle et al., 1973) in Finland
11      (Ojanpera et al., 1998). The following  12-h mean O3 exposures were observed in 1992:  14, 30,
12      61 ppb. In 1993, the values were 9, 21, and 45 ppb, (see Table 9-17 for AOT40 values and other
13      information). Yield was reduced 13% in each year by the added O3 treatment.
14          The effect of an intermittent constant addition of O3 using a free air exposure system was
15      investigated with the winter wheat cultivar Riband in Northumberland, UK (Ollerenshaw and
16      Lyons, 1999). Ozone exposures expressed as AOT40 values for September and October 1992
17      were 0.14 and 3.5 ppm-h; while for April to August 1993, values were 3.5 and 6.2 ppm-h. Yield
18      was reduced by 13%.
19          These results provide an additional line of evidence supporting the OTC-studies that
20      demonstrated yield reductions in wheat due to O3 exposures that occur in the United States.
21      These  yield reductions for wheat are generally similar to those reported previously in
22      22 comparable studies summarized in the 1996 AQCD (Table 5-25  of U.S. Environmental
23      Protection Agency, 1996).
24
25      9.6.4.2  Effects on Plant Quality
26          In addition to reductions in biomass or crop yield, O3 may also reduce the quality or
27      nutritive value of annual species. Many studies have shown effects of O3 on various measures of
28      plant organs that affect quality, with most studies focusing on characteristics important for food
29      or fodder (U.S. Environmental Protection Agency, 1996).
30          The effect of a continuous intermittent addition of O3 using a free air exposure system in
31      Northumberland, UK was investigated with the oilseed rape cultivar Eurol as discussed above

        January 2005                             9-217       DRAFT-DO NOT QUOTE OR CITE

-------
 1      (Ollerenshaw et al., 1999).  Ozone exposures expressed as AOT40 values for August to October
 2      1991 were 0.2 and 3.8 ppm-h and for June 1992 were 0.7 and 8.1 ppm-h. Yield quality measured
 3      as crude protein and oil content was decreased significantly. Because the price of the product is
 4      reduced in direct proportion to the oil content, such a decrease represents a substantial loss to
 5      growers (Ollerenshaw et al., 1999).
 6           Two wheat cultivars, Massey and Saluda, were each grown for one year each in Beltsville,
 7      MD (Table 9-17) and exposed to either CF or an addition of 40 ppb for 7 h/day for 5  days/week
 8      (Mulchi et al., 1995, 1996a; Rudorff et al.,  1996c). Milling and baking quality scores and flour
 9      protein were not significantly affected by elevated O3 exposure, but the softness equivalent was
10      increased slightly (2.4%) in both experiments (Rudorff et al., 1996b). The authors concluded
11      that these changes, along with other slight changes due to an increased CO2 treatment, suggested
12      that O3 and CO2 had only minor effects on wheat grain quality.  In wheat grown in Sweden, the
13      harvest index was significantly decreased and the protein content increased due to exposure to a
14      12-h mean of 48 ppb (Gelang et al., 2000).  In an analysis of 16 experiments conducted with
15      spring wheat and either O3 or CO2 exposures in four Nordic countries, a negative linear
16      relationship was found between grain yield and grain protein content (y = -0.38* +138.6,
17      expressed as percentages of the NF treatment (Pleijel et al., 1999a).
18           For three soybean cultivars grown in Raleigh, NC, O3 significantly decreased oleic acid
19      content, although the authors stated that the reduction was not large enough to be economically
20      important (Heagle et al., 1998).
21           In a UK study, potato exposed during 1998 to an AOT40 value of 12.5 ppm-h in OTCs
22      (in Nottingham) resulted in the paste from tubers being more viscous (Donnelly et al., 2001b).
23      In this study, an AOT40 exposure of 27.11  ppm-h in 1999 caused starch granules to be less
24      resistant to swelling, and total glycoalkaloid content was increased due to an increase in a-
25      solanine (Donnelly et al., 200Ib). Such increases in glycoalkaloid content have been observed
26      previously in potato (Pell and Pearson, 1984) and may be important, because  glycoalkaloids
27      cause bitter flavors and, at higher concentrations, cause toxicity. The authors indicated that
28      levels found  in this study approached those that may cause bitterness, but not those of concern
29      for toxicity (Donnelly et al., 2001b).
30           In the CHIP program the effects of O3 were studied using OTCs at six sites in northern
31      Europe, and yield decreases were observed as described above.  The reducing sugar and starch

        January 2005                             9-218       DRAFT-DO NOT QUOTE OR CITE

-------
 1     content of tubers decreased linearly due to O3 exposure, while the ascorbic acid concentration
 2     increased linearly (Vorne et al., 2002). Compared to the CF treatment, exposure to an AOT40
 3     value of 14 ppm-h decreased starch concentrations by 2%, decreased reducing sugar
 4     concentration by 30%, and increased ascorbic acid concentration by 20%. While the changes in
 5     reducing sugars and ascorbic acid increase tuber quality, the reduction in starch concentration
 6     decreases tuber quality.
 7          In two 1-year studies using OTCs in commercial fields in Spain, the soluble solids content
 8     of watermelon was decreased 4 to 8% due to seasonal O3 exposures as follows: AOT40 = 5.96
 9     ppm-h and SUM06 = 0.295 ppm-h in one year; and AOT40 = 18.9, SUM06 = 4.95 in the second
10     year (Gimeno et al., 1999).
11
12     9.6.4.3 Effects on Foliar Symptoms
13          For most annual crop species, the most important effects of O3 are on yield of the
14     commercially important part of the crop, expressed as the mass of the harvested portion.
15     However, for some crops, foliar symptoms are important if they reduce the marketability of the
16     crop. This is why efforts have been made to identify O3 exposures associated with foliar
17     symptoms. In Europe, Level I critical levels have been determined for such effects based on
18     observations from experiments conducted in 15 countries under the auspices of the United
19     Nations Economic Commission for Europe International Cooperative Programme on effects of
20     air pollution and other stresses on crops and non-woody plants (UN/ECE ICP-Vegetation;
21     formerly ICP-Crops), as well as on observations of symptoms in commercial fields from 1993  to
22     1996 (Benton et al., 1995, 2000).  Because the occurrence of symptoms increased with greater
23     humidity, these levels took into account the vapor pressure deficit (VPD). Two short-term
24     critical levels were derived from 1995 data:  an AOT40 value of 0.2 ppm-h over 5 days when
25     mean VPD is below 1.5 KPa (0930 - 1630 h), and a value of 0.5 ppm-h when the mean VPD is
26     above  1.5 Kpa (Benton et al.,  1996). The 1996 data  supported the critical levels in 83% of
27     observations, although symptoms occurred on three occasions when the AOT40 was less than
28     0.05 ppm-h and the VPD was very low — less than 0.6 Kpa.  The authors concluded that these
29     critical levels are a good indicator of the likelihood of foliar symptoms,  but that further
30     refinement may be required, such as including factors that modify O3  uptake by stomata.
       January 2005                             9-219        DRAFT-DO NOT QUOTE OR CITE

-------
 1           In a study in Germany, 25 native herbaceous species were exposed to several square-wave
 2      O3 exposures in CF OTCs (Bergmann et al., 1999).  Six of the 25 species showed O3-specific
 3      symptoms, and five species responded to single-day peaks of a single day.  The most sensitive
 4      species exhibiting O3-specific symptoms were Cirsium arvense and Sonchus asper, which
 5      responded to AOT40 values < 1.5 ppm-h (Bergmann et al., 1999).
 6           In the United States, attention has been paid to foliar symptoms on annual plants in
 7      previous criteria documents, and most recent literature has focused on other topics. However, a
 8      study of O3 effects was undertaken from 1990 to 1993 in Acadia National Park in Maine, which
 9      experiences elevated O3 exposures due to transport from urban areas located upwind (Kohut
10      et al., 2000). Because this study examined both herbaceous perennial species and woody
11      perennial species, the results are discussed in the perennial section below.  Other studies and
12      biomonitoring programs in other U.S. National Parks and over larger areas are oriented primarily
13      towards forested ecosystems, so such studies are also discussed in the perennial subsection
14      below as well as in Section 9.7.
15
16      9.6.4.4  Other Effects
17           In addition to the effects seen on individual species grown in monocultures, effects may
18      occur due to competition when species are grown in mixtures, as reviewed by Davison and
19      Barnes (1998).  Such effects have been investigated primarily in species grown for forage, which
20      often include perennial species;  therefore, this topic is discussed in the herbaceous perennial
21      section, Section 9.6.5.1.  Other stresses may interact with O3, including pathogens and pests as
22      well as environmental conditions such as drought.  These issues are discussed in Section 9.4.
23      Effects on reproduction are discussed briefly below.
24           Several studies during recent decades have demonstrated O3 effects on different stages of
25      reproduction. Effects of O3 have been observed on pollen germination, pollen tube growth,
26      fertilization, and abortion of reproductive  structures as reviewed by Black et al. (2000). This
27      issue is not addressed here because reproductive effects will  culminate for seed-bearing plants in
28      seed production, and the substantial body  of evidence relating O3 exposure and reduced seed
29      production is discussed above. However,  one example of a native species will be presented
30      because of its implications for extrapolating exposure-response data to noncommercial species.
31      Spreading dogbane has been identified as  a useful species for O3 biomonitoring because of

        January 2005                              9-220       DRAFT-DO NOT QUOTE OR CITE

-------
 1      O3-induced diagnostic symptoms (Kohut et al., 2000).  A study in Massachusetts found that
 2      exposure to O3 in NF OTCs or ambient plots for 103 days produced significantly fewer flowers,
 3      and also that fewer of these flowers survived to produce mature fruits (Bergweiler and Manning,
 4      1999). Because foliar symptoms were not common, the authors concluded they are not required
 5      for effects on reproduction to occur.  Genotoxic effects and effects  on population genetics are
 6      discussed in Section 9.4.
 7
 8      9.6.4.5  Scaling Experimental Data to Field Conditions
 9           Substantial effort has been invested in the design of OTCs for assessing the effects of air
10      pollutants on vegetation under near-ambient conditions. The design, construction, and
11      performance of many types of chambers has been reviewed extensively (Hogsett et al.,  1987a,
12      1987b).  Despite such design efforts, the influence of experimental  chambers on exposure-
13      response functions has been debated for many years (e.g., Manning and Krupa, 1992) because
14      several factors differ between OTC studies and actual fields. This issue is addressed in Section
15      9.2, and  only a few comments about the implications of chamber artifacts for interpreting
16      exposure-response relationships are presented here.
17           While it  is clear that chambers can alter some aspects of plant growth, the more important
18      issue is whether they alter the response of plants to O3. A review of such chamber studies done
19      in California found that plants responded similarly to O3 whether OTCs, closed-top chambers, or
20      air exclusion systems were used; differences were found for fewer than 10% of growth
21      parameters (Olszyk et al., 1986). In a different review of literature about Heagle-type OTCs
22      (Heagle  et al.,  1988), the authors concluded that "Although chamber effects on yield are
23      common, there are no results showing that this will result in a changed yield response to O3."
24      A more recent study of chamber effects examined the responses of tolerant and sensitive white
25      clover clones to ambient O3 in greenhouse, open-top, and ambient plots (Heagle et al., 1996).
26      For individual harvests, O3 reduced the forage weight of the sensitive clone 7 to 23% more in the
27      greenhouse than in OTCs. However, the response in OTCs was the same as in ambient plots.
28      Several studies have shown very similar yield response to O3 for plants grown in pots or in the
29      ground, suggesting that even such a significant change in environment does not alter the
30      proportional response to O3, at least as long as the plants are well watered (Heagle, 1979; Heagle
31      etal., 1983).

        January 2005                              9-221        DRAFT-DO NOT QUOTE OR CITE

-------
 1           Most experiments investigating the O3 effects on annual vegetation provide adequate water
 2      to avoid substantive drought stress. Because drought stress has generally been shown to reduce
 3      the effect of O3 on annual vegetation, such experiments may tend to overestimate O3 effects on
 4      crops and especially on unmanaged or seminatural vegetation.
 5           As mentioned above, the use of O3 flux rather than exposure is theoretically more realistic,
 6      and such an approach would also address the vertical gradient issue (Section 9.2). Recently, a
 7      number of investigators have suggested that modeling O3 flux can improve estimates of O3
 8      effects on vegetation.  Models of O3 flux can reduce the variation in the response to O3 that is
 9      sometimes observed between years in an experiment (Emberson et al., 2000a, 2000b; Fuhrer
10      et al., 1992; Griinhage et al., 1993; Griinhage and Haenel, 1997; Pleijel et al., 2000a). In a study
11      of O3 deposition to an oat crop in OTCs, O3 flux in the chamber was estimated to be up to twice
12      that in an adjacent field based on a K-theory approach and measurements of stomatal
13      conductance and environmental conditions (Pleijel et al., 1994b). These measurements were
14      made for 2 hours on 5 days when the canopy was physiologically active and wind speeds were
15      moderate. However, the O3 flux in a chamber without plants was nearly as high as that in the
16      open field. The authors conclude that O3 uptake in the chamber was between 100 and 200% of
17      that in the field. These models of flux have a sound biological and meteorological basis and are
18      useful for interpreting experimental data. Flux models have been successfully applied at
19      intensive study sites with detailed site-specific data on stomatal conductance and
20      micrometeorological conditions (e.g., Griinhage et al., 1993b, 1994; Fredericksen et al.,  1996).
21      Yet even at a single well-studied site, different methods can provide different estimates of O3
22      flux.  For example, at a site in a vineyard in California, an evapotranspiration-based method
23      overestimated the O3 flux as compared to an eddy covariance approach by 20 to 26% (Massman
24      and Grantz, 1995). At a site in a nearby cotton field the evapotranspiration-based approach
25      overestimated the eddy-covariance method by 8 to 38%.
26           Interest has been increasing in recent years, particularly in Europe, in using mathematically
27      tractable flux models for O3 assessments at the regional and national scale (Emberson et al.,
28      2000a, 2000b). However, methods for scaling site-specific models of O3 flux to large areas
29      remains an active and challenging area of research (Massman et al., 1994; Massman and Ham,
30      1994; Wesely and Hicks, 2000).  Reducing uncertainties in flux estimates for areas with diverse
31      surface or terrain conditions to within ±50% requires "very careful application of dry deposition

        January 2005                             9-222        DRAFT-DO NOT QUOTE OR CITE

-------
 1      models, some model development, and support by experimental observations" (Wesely and
 2      Hicks, 2000). As an example, the annual average deposition velocity of O3 among three nearby
 3      sites in similar vegetation was found to vary by ±10%, presumably due to terrain (Brook et al.,
 4      1997). Moreover, the authors stated that the actual variation was even greater, because stomatal
 5      uptake was unrealistically assumed to be the same among all sites, and flux is strongly
 6      influenced by stomatal conductance (Brook et al., 1997).  Stomatal conductance is affected by
 7      factors such as temperature, vapor pressure deficit, plant water status, and temperature. It is a
 8      challenging task to obtain such data for regional, national, or continental flux modeling (Pleijel,
 9      1998). This topic is addressed further in Section 9.5.
10           If a flux approach is not used, another issue in scaling from experimental data to national
11      assessments is the selection of an O3-exposure index. As discussed in Section 9.5, there is no
12      evidence to support a single value for defining a  cutoff value for calculating a peak-weighted O3
13      index. In Europe, it has become common practice to use a cutoff value of 40 ppb, as in the
14      AOT40 index; while in the United States, a cutoff value of 60  ppb has been used, as in the
15      SUM06 index. The W126 index has been promoted, because it has a continuous weighting
16      function without an arbitrary cutoff value; but few exposure-response studies have used this
17      index to date. There is evidence from some European studies that a cutoff value either lower
18      (Pleijel et al., 1997) or higher (Finnan et al., 1996, 1997) may  provide a better statistical fit to
19      experimental data. However, the choice of the type of weighting depends on the type of
20      response model that is fit. In an evaluation of numerous indices fit to seven studies of the
21      response of spring wheat to O3 in OTCs in Northern Europe, different types of indices performed
22      better for linear response models than for Wiebull response models (Finnan et al., 1997).  In this
23      study, cumulative indices that give greater weight to higher O3 concentrations performed best;
24      and, while the AOT40 was not the best index, it did perform well, and slightly better than the
25      SUM06 in conjunction with a linear model (Finnan et al., 1997). Although this section is
26      focused on annual plants, the issue of cutoff values applies to perennial species as well;
27      therefore, two examples of tree species will be discussed here. Aspen seedlings were exposed to
28      O3 in controlled environment chambers for 4 weeks, and a significantly greater degree of foliar
29      symptoms occurred with a pattern of variable peak exposures compared to a constant peak
30      exposure with the same total SUM06 exposure (Yun and Laurence,  1999b). However, a field
31      experiment in large OTCs in Ithaca, NY with these two clones exposed to ambient, 1.7*,- and

        January 2005                             9-223        DRAFT-DO NOT QUOTE OR CITE

-------
 1      3 x-ambient O3 concentrations (SUM06 = 1, 20, and 62 ppb) found SUM06 to be highly
 2      correlated with foliar symptoms and biomass effects (Yun and Laurence, 1999a). When silver
 3      birch was exposed to different patterns of O3 with the same AOT40 value in growth chambers,
 4      high peak concentrations and exposure shape were found to be important for symptom
 5      production (Oksanen and Holopainen, 2001). However, growth reductions were best predicted
 6      by total cumulative exposure.
 7           An investigation of different exposure indices was conducted using data from experiments
 8      conducted during 1993 at eight sites in the eastern United States in which O3 sensitive and
 9      O3-tolerant white clover genotypes were grown using methods developed by Heagle  et al.
10      (1994). A statistical approach based on profile likelihoods was used to estimate parameters in
11      generalized exposure indices similar to the  SUM06 and AOT40 indices (Blankenship and
12      Stefanski, 2001). The results showed that for the  SUMX family of indices, where X is a cutoff
13      value, hourly O3 concentrations over -71 ppb contribute the most to yield prediction. For the
14      AOTX family of indices, the parameter was 54.4.  These values are similar to those used in the
15      SUM06 and AOT40 indices already in use. Furthermore, investigation of weighting for time of
16      day confirmed the importance of the mid-afternoon hours for this data set, unlike the results
17      found for wheat in Sweden (Danielsson et al., 2003; Pleijel et al.,  2000a).
18
19      9.6.4.6 European Critical Levels
20           During the late 1980s and 1990s, substantial effort was expended in Europe to quantify the
21      effects of O3 on crops and trees. A focus for this research effort was to develop and refine
22      "critical levels" for O3 under the auspices of the United Nations Economic Commission for
23      Europe (UNECE) Convention on Long-Range Transboundary Air Pollution (Sanders et al.,
24      1995). A critical level was defined as "the  concentration of pollutant in the atmosphere above
25      which direct adverse effects on receptors , such as plants, ecosystems or materials, may occur
26      according to present knowledge" (United Nations  Economic Commission for Europe, 1988).
27      Areas where the levels were exceeded were identified to plan abatement strategies.  Critical
28      levels are not air quality standards as  are used in the United States for criteria pollutants, nor are
29      they legally enforced. However, they have been used successfully for planning reductions in
30      sulfur and nitrogen pollution. Critical levels for O3 are intended to prevent long-term deleterious
31      effects on the most sensitive plant species under the most sensitive environmental conditions, but

        January 2005                             9-224       DRAFT-DO NOT QUOTE OR CITE

-------
 1      not to quantify O3 effects. The nature of the "significant harmful effects" is not specified in the
 2      original definition, which allows different levels to be defined for different effects, such as foliar
 3      symptoms or reduction in crop yield. Different levels also have been set for crops, forests and
 4      seminatural vegetation.  The caveat, "according to present knowledge" is important, because
 5      critical levels are revised periodically as new scientific information becomes available.
 6           Because these critical levels were designed for use in regional modeling and mapping, they
 7      were required to be of a simple form (Sanders et al.,  1995). During the early 1990s, the AOT40
 8      index was selected as the form of the critical level, and Level I critical levels were determined
 9      for arable crops and forestry.  The AOT40 index is defined as the  sum of the difference between
10      the hourly concentration (in ppb) and 40 ppb when the concentration exceeds 40 ppb for the
11      hours when global radiation exceeds 50 W nT2 (Fuhrer et al., 1997). In regression analysis of
12      15 OTC studies of spring wheat, including one study from the United States and 14 from
13      locations ranging from southern Sweden to  Switzerland, an AOT40 value of 5.7 ppm-h was
14      found to correspond to a 10% yield loss, and a value of 2.8 ppm-h corresponded to a 5% yield
15      loss (Fuhrer et al., 1997). Because a 4 to 5% yield loss could be detected with a confidence level
16      of 99%, a critical level of 3 ppm-h was selected an AOT40 value of in 1996 (Karenlampi and
17      Skarby, 1996). This critical level is defined for a 3-month period  calculated for daylight hours.
18      It was suggested that a 5-year average should be used to assess the exceedance of this critical
19      level, because this value was associated with a small yield decrement (Fuhrer et al., 1997). This
20      value is currently used for all crops, because it is the best-supported value and because the
21      limited data from other crop species do not provide strong evidence that a more stringent value is
22      required (Fuhrer et al., 1997).
23           The Level I critical levels were successfully used in the 1990s to map areas of exceedance,
24      but the research led to the conclusion that simple, exposure-based levels lead to overestimation
25      of the effects in some regions and underestimation in others.  The main problem is that other
26      environmental  factors (e.g., vapor pressure deficit, water stress, temperature, light) alter O3
27      uptake  and its effects. Therefore, the decision was made to work towards a flux-based approach,
28      aiming to be able to model O3  flux-effect relationships for the three vegetation types (i.e.,  crops,
29      forests, seminatural vegetation).  Progress has been made in modelling flux under experimental
30      conditions (Danielsson et al., 2003; Griinhage et al.,  2001; Griinhage and Jager, 2003; Pleijel
        January 2005                             9-225        DRAFT-DO NOT QUOTE OR CITE

-------
 1      et al., 2000a), but because the UNECE meeting in November 2002 concluded that more research
 2      is needed, the subject is to be reassessed in 2 to 5 years' time.
 3           There has been criticism that the Level I critical level for crops overestimates O3 effects in
 4      the Mediterranean countries, because it was developed based on studies in Northern Europe
 5      (De Santis, 1999, 2000).  However, there is evidence of substantial crop loss due to O3 in some
 6      southern European counties, such as the Po valley in Northern Italy (Fumagalli et al., 2001).
 7      In these studies Heagle type OTCs were used. Losses in NF chambers as compared to CF
 8      chambers over several years at two sites ranged from 11.2 to 22.8% for barley and wheat, from
 9      0.3 to 31.5% for other crop  species, and from 4.1 to 19.8% for forage species (Fumagalli et al.,
10      2001). Surprisingly, the least effect was observed for soybean, despite AOT40 values of 9.32
11      ppm-h, 3 x the Level I critical level.  Similarly, a review of studies in  Northern Italy found that
12      ambient O3 episodes have been reported to cause foliar symptoms on 24 agricultural and
13      horticultural crops in commercial fields (Fumagalli et al., 2001). Ambient O3 has also been
14      reported to cause yield losses in several crop species, although no data on O3 exposure were
15      presented by the reviewers (Fumagalli et al., 2001).
16           The Level I approach  has also been criticized for focusing only on a single annual species
17      (wheat) and a single woody perennial species (beech).  However, this species focus is
18      appropriate, because the goal was to determine an exposure-response relationship for a sensitive
19      species based on available data.  In support of standards in Germany, an effort was made to
20      combine data from different species, and consisted of a meta-analysis of studies conducted in
21      both closed chambers and OTCs (Griinhage et al., 2001).  In this study, experiments published
22      between 1989 and 1999 were analyzed if they met the following conditions:  (1) a significant O3
23      effect was determined (2) exposure conditions were well defined, (3) foliar symptoms, growth,
24      or yield was measured, and  (4) plant species were relevant to Europe (Griinhage et al., 2001).
25      Despite the focus on European species, many of the species studies also occur in the United
26      States. Separate regressions for herbaceous plants and for tree species were created as a function
27      of duration of exposure at a given level of O3 exposure at the top of the plant canopy.  These
28      regression equations, with confidence limits and with correction for the vertical gradient in O3
29      from the top of the quasi-laminar boundary layer, can be used to define whether effects are
30      unlikely  (below the lower confidence limit), probable (between the confidence limits), or highly
31      likely (above the upper confidence limit) to occur near a given O3-monitoring station.

        January 2005                              9-226        DRAFT-DO NOT QUOTE OR CITE

-------
 1          A further concern about the Level 1 approach is that foliar symptoms, rather than biomass,
 2     may be an important endpoint, because foliar symptoms may be more sensitive than is biomass
 3     increment (VanderHeyden et al., 2001). In an OTC study in southern Switzerland, it was shown
 4     that a number of tree species show foliar symptoms at AOT40 values lower than the Level 1
 5     value of 10 ppm-h (VanderHeyden et al., 2001).
 6
 7     9.6.4.7  Summary of Effects on Short-Lived Species
 8          For annual vegetation, the data summarized in Table 9-17 show a range of growth and
 9     yield responses both within species and among species. Nearly all of these data were derived
10     from studies in OTCs, with only two studies using open-air systems in the UK (Ollerenshaw
11     et al., 1999; Ollerenshaw and Lyons,  1999). It is difficult to compare studies that report O3
12     exposure in different indices such as AOT40, SUM06, or 7-h or 12-h mean values. However,
13     when such comparisons can be made, the results of this recent body of research confirm earlier
14     results summarized in the 1995 AQCD (U.S. Environmental Protection Agency, 1996).
15     A summary of earlier literature concluded that a 7-h, 3-month mean of 49 ppb  corresponding to a
16     SUM06 exposure of 24.4 ppm-h would cause 10% loss in 50% of 49 experimental cases (Tingey
17     et al., 1991). A similar study using a 24-h, rather than 7-h, averaging period found that a
18     SUM06 exposure of 26.4 ppb would cause 10% loss in 50% of 54 experimental cases (Lee et al.,
19     1994a, 1994b).  Recent data summarized in Table 9-17 support this conclusion. These values
20     represent ambient exposure patterns that occur in some years over large portions of the United
21     States.  Some annual species such as soybean are more sensitive, and greater losses may be
22     expected (Table 9-17). Thus, the recent scientific literature supports the conclusions of the 1996
23     AQCD that ambient O3 concentrations are reducing the yield of major crops in the United States.
24          Much research in Europe has used the AOT40 exposure statistic, and substantial effort has
25     gone into developing Level-1 values for vegetation.  Based on regression analysis of 15 OTC
26     studies of spring wheat, including one study from the United States and 14 from locations
27     ranging from southern Sweden to Switzerland, an AOT40 value of 5.7 ppm-h was found to
28     correspond to a 10% yield loss, and a value of 2.8 ppm-h corresponded to a 5% yield loss
29     (Fuhrer et al., 1997).  Because a 4 to 5% decrease could be detected with a confidence level of
30     99%, a critical level of an AOT40 value of 3 ppm-h was selected in 1996 (Karenlampi and
31     Skarby, 1996).

       January 2005                             9-227       DRAFT-DO NOT QUOTE OR CITE

-------
 1
 2
 3
 4
 5
 6
 7
 8
 9
10
11
12
13
14
15
16
     In addition to reductions in crop yield, O3 may also reduce the quality or nutritive value of
annual species.  Many studies have shown effects of O3 on various measures of plant organs that
affect quality, with most studies focusing on characteristics important for food or fodder. These
studies indicate that there may be economically important effects of ambient O3 on the quality of
crop and forage species. Previous criteria documents have concluded that foliar symptoms on
marketable portions of crops and ornamental plants can occur with seasonal 7-h mean O3
exposures of 40 to 100 ppb (U.S. Environmental Protection Agency,  1978, 1986, 1996).  The
recent scientific literature does not refute this conclusion.
     The use of OTCs may reverse the usual vertical gradient in O3 that occurs within a few
meters above the ground surface (Section 9.2).  Such a reversal suggests that OTC studies may
overestimate, to some degree, the effects of an O3 concentration measured several meters above
the ground. However such considerations do not invalidate the conclusion of the 1996 AQCD
(U.S. Environmental Protection Agency, 1996) that ambient (Table 9-14, Table 9-21) O3
concentrations are sufficient to reduce the yield of major crops in the United States.
            Table 9-21. Ozone Exposures at 35 Rural Sites in the Clean Air Status and Trends
                  Network in the Central and Eastern United States From 1989 to 1995
Subregion
Midwest
Upper
Midwest
Northeast
Upper
Northeast
South
Central
Southern
Periphery
SUM06
12-h, 3-Month
Mean
31.5
18.9
33.2
13.2
34.5

19.2

UM06
12-h,
3-Month SD
10.2
8.5
11.9
8.6
16.6

7.6

W126
3-Month
Mean
25.1
16.0
26.6
12.8
25.6

15.2

W126
3-Month
SD
7.7
5.9
9.5
6.5
11.5

5.4

Max.
8h > 80 ppb(n)
Mean
13.8
5.6
15.8
3.3
7.1

1.9

Max.
8h > 80 ppb(n)
SD
10.6
5.6
12.2
5.5
10.0

1.6

        Units for SUM06 and W126 are ppm-h.
        Source: Baumgardner and Edgerton (1998).
       January 2005
                                         9-228
DRAFT-DO NOT QUOTE OR CITE

-------
 1      9.6.5   Effects of Ozone on Long-Lived (Perennial) Species
 2          Although there has been considerable research in Europe on annual species during the past
 3      10 years, much recent research in the United States has focused on perennial species.  In Europe,
 4      and in a few studies in the United States, effects of O3 on mixtures of annual and perennial
 5      herbaceous species have been investigated using growth chambers, greenhouses, and OTCs.
 6      Section 9.6.5.1 reviews such studies, with an emphasis on studies using OTCs.
 7
 8      9.6.5.1 Herbaceous Perennial Species
 9          Two alfalfa cultivars were grown in pots and exposed to CF, NF, 1.5 x -ambient and
10      2x-ambient O3 concentrations in two 1-year studies in Quebec, Canada (Renaud et al., 1997).
11      One cultivar, Apica, is commonly grown in the region, and another, Team, is normally grown
12      farther south and is more tolerant to O3. For Apica in both years and for Team in 1991, O3
13      exposure caused a linear reduction  in above-ground biomass.  In the NF treatment, growth of
14      Apica  was decreased by 31 and 21% in the 2 years, while the growth of Team was reduced by
15      14% in 1991, but not reduced in 1992. The authors suggested that the differing effects on Team
16      could be due to different progenies and propagation methods in the 2 years or to more rapid
17      growth in 1991 along with higher O3 peak values in 1991.  In 1991, O3 maxima exceeded 60  ppb
18      in 15 days, whereas in 1992 there were only three such days. At the end of the growing season,
19      total starch reserves in roots were decreased by O3 due primarily to a decrease in root mass,
20      which  the authors suggested could  accelerate decline in alfalfa yields. These yield reductions are
21      generally similar to those reported previously in five  similar studies summarized in the 1996
22      AQCD (Table 9-25 of U.S. Environmental Protection Agency, 1996).
23          A study in Alabama exposed  early- and late-season-planted bahia grass (cultivar
24      Pensacola) in OTCs to CF, NF or 2x-ambient O3 treatments (Muntifering et al., 2000). Ozone
25      exposures expressed as 12-h mean values over the 24-week experiment were 22, 45, and 91 ppb,
26      and the highest ambient O3 concentrations were recorded in late June, late July, late August and
27      mid-September at approximately 90 ppb.  Above-ground biomass growth was reduced by the NF
28      treatment for the first and second harvest by 34% and 29% for the early-season planting, but
29      statistically significant effects were not observed in the late-season planting (Table 9-18).  The
30      2x-ambient treatment did not cause further significant reductions in biomass.  The authors
31      suggested that the lack of a significant O3 effect in the late planting may have been due to the

        January 2005                             9-229       DRAFT-DO NOT QUOTE OR CITE

-------
 1      shorter total O3 exposure time as well as to the lower O3-exposure concentrations during the
 2      weeks immediately preceding harvest. These results are important, because this is an
 3      economically important species and because previous studies have focused on grass species that
 4      use the C3, rather than C4, metabolic pathway.
 5           An investigation of the use of different O3 indices and averaging times on the correlation
 6      with growth effects was undertaken with the NC clover system (Heagle et al., 1995). For 2 years
 7      of data at six sites in Massachusetts, Oregon, North Carolina, California (2 sites), and Virginia,
 8      averaging time was found to be more  important than the choice of the type of index including
 9      mean, SUM06, and AOT40 (Heagle and Stefanski, 2000).  The best correlation between O3
10      exposure and the ratio of sensitive to tolerant clover types was found for the 6-h period from
11      1000 to 1600 h.  For this period, very  similar r2 values (0.91 to 0.94) were found for SUM06,
12      W126, and AOT40 (Heagle and Stefanski, 2000).  For the above indices, a linear relationship
13      was found, with no effect in Corvallis, OR with exposure to a SUM06 value of 10.2 ppm-h and a
14      ratio of 0.53 (sensitive/tolerant) at San Bernardino with a SUM06 exposure of 39.4 ppm-h
15      (Heagle and Stefanski, 2000).
16           In a study of the biomass ratio of O3-sensitive versus O3-insensitive clover at 14 sites in
17      Europe during 1996 to 1998, a model  that was developed using artificial neural network (ANN,
18      see Section 9.2) techniques had r2 values for the training data of 0.84 and for unseen validation
19      data of 0.71 (Mills et al., 2000). The predictive factors in the model were AOT40, 24-h mean
20      O3, daylight mean temperature, and 24-h mean temperature. This model was selected after a
21      thorough investigation of a number of models using many more or fewer parameters using both
22      ANN and multiple linear regression techniques. This model predicted that a 5% reduction in
23      biomass ratio was associated with AOT40 values in the range 0.9 to 1.7 ppm-h accumulated over
24      28 days, with plants being most sensitive under conditions  of low NOX, moderate temperature,
25      and high 24-h mean O3 concentration.
26           Two experiments in Ontario investigated effects of square-wave additions of O3 for 7 h on
27      1  day/week for 7 weeks on the growth and fruit production of two cultivars of primocane (first-
28      year fruiting) raspberry plants (Sullivan et al., 1994). Plants were grown in OTCs and exposed
29      to ambient O3 (mean values of < 4 ppb). In the both experiments, additions of 12 ppb O3 on
30      1  day/week for 7 weeks did not significantly reduce fruit yield. In the second experiment,
        January 2005                             9-230        DRAFT-DO NOT QUOTE OR CITE

-------
 1      additions of 24 ppb significantly reduced the fruit yield 52% below that of the ambient treatment
 2      in one cultivar (Heritage), but not in another (Redwing).
 3           Exposure to a square-wave 8-h mean O3 concentration of 92 ppb for 62 days in an
 4      experiment in OTCs the UK did not significantly reduce the total yield of strawberry fruits, but
 5      did decrease the average size of the fruits by 14% (Drogoudi and Ashmore, 2000).  This
 6      contrasts with an increase in total yield (fruit weight) found in a previous study in California
 7      (Takemotoetal., 1988).
 8           When timothy was exposed in OTCs in Sweden to NF, CF, and CF+O3 treatments, there
 9      was no effect of a 12-h mean O3 exposure of 68 ppb (NF treatment), but a 12-h mean exposure
10      of 152 ppb decreased yield by 58% (Danielsson et al.,  1999).  A similar lack of effect of
11      exposure to a 12-h mean O3 exposure of 62 ppb was found in a previous study in the United
12      States (Kohut et al., 1988).
13           Although most investigations of O3-response relationships focus on growth or yield of
14      marketable portions of plants, some studies also investigate effects on plant  quality. In the study
15      of bahia  grass in Alabama discussed above, in addition to the effects on yield, there were
16      significant effects on quality for ruminant nutrition (Muntifering et al., 2000). Concentrations of
17      neutral detergent fiber (NDF) were higher in primary-growth and regrowth forages from the
18      early-season planting when exposed to 2x-ambient O3  than when exposed to the NF treatment.
19      The concentration of acid detergent fiber was higher in the 2x-ambient treatment than in NF
20      treatment regrowth, whereas acid detergent lignin concentration was higher  in 2 x-ambient than
21      in NF primary-growth forage. Crude protein concentrations were lower in CF-exposed than in
22      NF-exposed regrowth forage from the early planting and in CF- than in NF-exposed primary -
23      growth forage from the initial harvest of the late-season planting.  No differences were observed
24      among treatments in concentrations of total phenolics in primary-growth or regrowth forages
25      from either planting, although concentrations of total phenolics tended to be higher in
26      CF-exposed than in NF-exposed primary-growth forage from the late-season planting. The
27      authors concluded that the alterations in quality of primary-growth and vegetative regrowth
28      forages were of sufficient magnitude to have nutritional and possibly economic implications to
29      their use for ruminant animal feed.
30           Sericea lespedeza and little bluestem were exposed to CF, NF, and 2 x-ambient O3 in OTCs
31      in Alabama for 10 weeks (Powell et al., 2003). Ozone treatments expressed as 12-h mean

        January 2005                             9-231       DRAFT-DO NOT QUOTE OR CITE

-------
 1      concentrations were 23, 40, and 83 ppb and expressed as seasonal SUM06 values were 0.2, 9.1,
 2      and 61.0 ppm-h. Although there were few statistically significant effects of O3 on yield (the
 3      yield of only the 2x-ambient compared to NF for Sericea lespedeza in the last of six harvests),
 4      plant quality as feed for ruminants was  reduced. The nutritive quality of Sericea lespedeza was
 5      decreased by 7% and that of little  bluestem by 2% as a result of increased cell wall constituents
 6      and decreased in vitro digestibility.
 7           For some annual species, particularly crops, the endpoint for an assessment of the risk of
 8      O3 exposure can be defined as yield or growth; e.g., production of grain.  For plants grown in
 9      mixtures such as hayfields, and natural  or seminatural grasslands (including native
10      nonagricultural species), endpoints other than production of biomass may be important.  Such
11      endpoints include biodiversity or species composition and measures of plant quality such as total
12      protein, and effects may result from competitive interactions among plants in mixed-species
13      communities. Most of the available data on non-crop herbaceous species are for grasslands.
14           In a study of two perennial grasses (bent grass and red fescue) and two forbs (white clover,
15      Germander speedwell) grown in pots in OTCs, O3 effects differed among species and cutting
16      treatments (Ashmore and Ainsworth, 1995) (see also Table 9-18). Fescue biomass increased
17      with higher O3 treatments both in pots that were not cut during the growing season (mid-June to
18      mid-September), and those that were  cut every two weeks.  However, bent grass biomass
19      decreased with higher O3 exposure in the uncut treatment and increased in the cut treatment.
20      White clover and Germander speedwell biomass decreased substantially with higher O3 exposure
21      with and without cutting, with greater decreases in the cut treatment.  The authors cautioned that
22      the experiment did not replicate field circumstances. The plants were all cut to  1 cm above the
23      ground, which does not simulate grazing, and there may have been effects due to growing the
24      plants in pots. However, two key  results of this study likely apply to mixtures of species
25      growing in hay or forage fields or seminatural and natural communities.  First, O3 exposure
26      increased the growth of O3-tolerant species while exacerbating the growth decrease of O3
27      sensitive species. Second, the total biomass of the mixed-species community was unaffected by
28      O3 exposure due to the differential effects on O3-sensitive and O3-tolerant species.
29           In a 2-year study using OTCs placed over managed pasture in Switzerland, the above-
30      ground biomass of clover (red and white) was reduced linearly in response to increased O3
31      exposure (Table 9-18) (Fuhrer et al., 1994). Exposure to a 12-h mean concentration of 39 ppb

        January 2005                              9-232        DRAFT-DO NOT QUOTE OR CITE

-------
 1      O3 reduced biomass by 24% as compared to the CF treatment with a 12-h mean concentration of
 2      21 ppb O3. There was a trend toward increased above-ground biomass of grasses (primarily
 3      orchard grass), but this trend was not statistically significant. As often found in other studies of
 4      mixtures of species, by O3 exposure did not significantly affect total above-ground biomass O3
 5      exposure.
 6           A field-grown grass/clover mixture was exposed to CF, NF, and two O3 addition treatments
 7      for two growing seasons in OTCs in southern Sweden (Pleijel et al., 1996).  The mixture
 8      consisted of 15% (by seed weight) red clover cv. Fanny, 60% timothy cv. Alexander, and 25%
 9      fescue cv. Svalofs Sena.. Ozone concentrations expressed as AOT40 ranged from 0 to
10      approximately 47 ppm-h and expressed as  7-h mean from 11 to 62 ppb. Over this range, a slight,
11      but statistically significant, linear decrease of 4% in total above-ground biomass was seen
12      growth over six harvests. No  significant decrease was seen in the proportion of clover, and the
13      authors ascribed this lack of effect to the relatively higher O3 sensitivity of timothy and lower
14      sensitivity of this clover cultivar as compared to previously published results for other
15      grass/clover mixtures (e.g., Fuhrer et al., 1994).
16           A mixture of species in an old farm field in Alabama was exposed to O3 for two growing
17      seasons in large OTCs 4.8 m high, and 4.5 m  diameter); and a similar lack of effect of O3 was
18      found on total plant community growth measured as both canopy  cover and vertical canopy
19      density (Barbo et al., 1998). Of the 40 species in the plots, O3 effects were examined only on the
20      most common species:  blackberry, broomsedge bluestem, bahia grass, Panicum spp., and
21      winged sumac (second year only). Of these species, a 2x-ambient O3 treatment increased the
22      percent canopy cover of blackberry over 2 years by 124%, while that of winged sumac was
23      decreased by 95% (Table 9-18). Blackberry showed no significant effect on biomass, but ripe
24      fruit mass was decreased by 28% (Chappelka, 2002).  However, there was a significant chamber
25      effect for this latter response.  Biomass was not reported for other species in this study due to a
26      hurricane. Effects on loblolly pine grown  in this experiment are discussed subsequently in
27      Section 9.6.5.5.
28           In summary, results of studies on perennial herbaceous species conducted since the 1996
29      criteria document was prepared are presented in Table 9-18. As for single-season agricultural
30      crops, yields of multiple-year  forage crops are reduced at O3 exposures that  occur in some years
31      over large areas of the United States (Table 9-14, Table 9-21). This result confirms that reported

        January 2005                              9-233        DRAFT-DO NOT QUOTE OR CITE

-------
 1      in the 1996 AQCD (U.S. Environmental Protection Agency, 1996).  When species are grown in
 2      mixtures, O3 exposure can increase the growth of O3-tolerant species while exacerbating the
 3      growth decrease of O3-sensitive species (e.g., Ashmore and Ainsworth, 1995; Fuhrer et al.,
 4      1994). Because of this competitive interaction, the total growth of the mixed-species community
 5      may not be affected by O3 exposure (Ashmore and Ainsworth, 1995; Barbo et al., 1998; Fuhrer
 6      et al., 1994). However, in some cases mixtures of grasses and clover species have shown
 7      significant decreases in total biomass growth in response to O3 exposure in studies in the United
 8      States (Heagle et al., 1989; Kohut et al., 1988) and in Sweden  (Pleijel et al., 1996).  In Europe,
 9      a provisional critical level for perennials of an AOT40 value of 7 ppm-h over 6 months has been
10      proposed to protect sensitive plant species from the adverse effects of O3.
11
12      9.6.5.2   Deciduous Woody Species
13           It is extremely difficult and costly to study entire mature trees under controlled conditions
14      such as those in OTCs, with the possible exception of some species  managed for fruit or nut
15      production.  For this reason, the great maj ority of investigations have been of seedlings in
16      growth chambers, greenhouses, or OTCs.  A few investigations have been carried out on
17      saplings or more mature trees using free air exposure systems (Haeberle et al., 1999; Isebrands
18      et al., 2000, 2001; Werner and Fabian,  2002).  Exposure-response functions based on 28
19      experimental cases of seedling response to O3 suggest that a SUM06 exposure for 3 months of
20      31.5 ppm-h would protect hardwoods from a 10% growth loss  in 50% of the cases (Table 9-19).
21      However, there is a substantial range in sensitivity among species.  A risk analysis was
22      undertaken to predict tree biomass growth reductions due to O3 based on exposure-response
23      equations for seedlings of individual species combined with the species' spatial distribution
24      across the eastern United States and interpolated  O3 exposure expressed as SUM06 (Hogsett
25      et al., 1997). The growth of sensitive species such as aspen and black cherry was predicted to be
26      reduced by at least 20% across 50% of their ranges in a high O3 year and approximately 10% in a
27      lower-than-average O3  year (Hogsett et al., 1997).
28           A few investigations reported since the last criteria document was prepared have examined
29      saplings or mature trees, notably of oak species in the southern Appalachian Mountains and pine
30      species in California. Most of these studies have been of natural (uncontrolled) O3 exposures.
31      Additional studies have examined foliar symptoms on mature trees,  and in recent years such

        January 2005                              9-234      DRAFT-DO NOT QUOTE OR CITE

-------
 1      surveys have become more common and with greater attention to the standardization of methods
 2      and the use of reliable indicator species (Campbell et al., 2000;  Smith et al., 2003).  Previous
 3      criteria documents have noted the difficulty in relating foliar symptoms to effects on individual
 4      tree growth, stand growth, or ecosystem characteristics (U.S. Environmental Protection Agency,
 5      1996). This difficulty still remains to the present day.
 6           Some investigators have suggested that a comprehensive risk assessment of the effects of
 7      O3 on mature tree species might best be accomplished by extrapolating measured effects of O3
 8      on seedlings to effects on forests using models based on tree physiology and forest stand
 9      dynamics (Fuhrer et al., 1997; Hogsett et al., 1997; Chappelka,  1998; Laurence et al., 2000,
10      2001). Several such efforts are discussed in Sections 9.4 and 9.7.
11           In this subsection, emphasis will be placed on experimental evidence of O3 effects on the
12      growth of woody species under controlled conditions with some information from observational
13      studies under ambient conditions in forests.  Experimental results are summarized for deciduous
14      species in Table 9-20; and the species are discussed below in the order in which they appear in
15      this table.
16           A series of studies out in Michigan and Wisconsin during the 1990s on clones of trembling
17      aspen previously demonstrated that they differ in their O3 sensitivity (Coleman et al., 1995b,
18      1995a, 1996; Dickson et al., 2001; Isebrands et al., 2000, 2001; Karnosky et al.,  1996, 1998,
19      1999; King et al., 2001).  Several of those studies were undertaken with plants in pots or in the
20      ground in OTCs and additional studies were undertaken at three sites selected to differ primarily
21      in O3 exposure  (Karnosky  et al., 1999).  An ongoing study was undertaken using a free air
22      carbon dioxide and O3 enrichment (FACE) facility in Rhinelander, WI (Isebrands et al., 2000,
23      2001). These studies showed that O3-symptom expression was  generally similar in OTCs and
24      FACE and gradient sites, supporting the previously observed variation among aspen clones
25      (Karnosky et al., 1999). In the Michigan OTC studies, O3 decreased total plant biomass by an
26      average of 18% in each of 2 years for three clones classified as  high, intermediate and low in O3
27      tolerance (clones 216, 271 and 259; Karnosky et al. [1996]). However, in the first experiment,
28      an additional clone known to be O3 insensitive showed no response (clone 253). In these studies,
29      plants were grown in pots  and exposed to CF, 1 x-ambient, and 2x -ambient O3 treatments in two
30      separate experiments of 98 days each (additional treatments of 0.5x-ambient and 1.5x-ambient
31      were used in the first experiment  only (Karnosky et al., 1996).  Ozone concentrations expressed

        January 2005                             9-23 5        DRAFT-DO NOT QUOTE OR CITE

-------
 1     as 3-month, 7-h mean values for 1990 ranged from 7 to 69 ppb for the CF to 2x-ambient
 2     treatments; while for 1991, values ranged from 22 to 92 ppb.
 3          The FACE study evaluated the effects of 3 years of exposure to combinations of elevated
 4     CO2 and O3 on growth responses in mixture of five trembling aspen clones (Isebrands et al.,
 5     2000, 2001).  Height, diameter, and stem volume (diameter2 x height) were decreased by
 6     elevated O3. On average for all clones, stem volume was decreased by 20% over the 3 years in
 7     the elevated O3 treatment as compared with the 1 x-ambient treatment.  However, one clone
 8     showed increased growth in response to O3. Ozone concentrations were not reported. This
 9     FACE facility study is important, because it confirmed responses reported previously with these
10     clones grown in pots or soil in OTCs, without the alterations of microclimate induced by
11     chambers.  Currently, this is the only U.S. study using this technology to have examined the
12     effects of O3 under these conditions. This study is also significant, because the elevated
13     O3-exposure pattern used was intended to reproduce the 6-year average pattern from Washtenaw
14     County, Michigan (Karnosky et al., 1999).
15          Rooted cuttings of two aspen clones from Acadia National Park in Maine were exposed to
16     1 x-ambient, 1.7x-ambient, and 3x-ambient O3 concentrations in large OTCs in Ithaca, NY for
17     much of one growing season (15 June to 15 September) (Yun and Laurence, 1999a). Both
18     circular (4.7 m diameter, 3.7 m height) and rectangular (7.4 m x 2.75 m x 3.7 m height)
19     chambers were used (Mandl et al., 1989). Exposure to 1.7x-ambient O3 (SUM06 = 20 ppm-h,
20     9-h mean = 74 ppb) reduced shoot growth by 14 and 25% compared to ambient O3 for the two
21     clones (Yun and Laurence, 1999a). Total dry weight was reduced by 55 and 35% in the two
22     clones by the 3 x-ambient treatment (SUM06 = 62 ppm-h, 9-h mean =124 ppb) compared to the
23     ambient O3 treatment.
24          When black poplar cuttings in OTCs in Belgium were exposed to 8-h mean O3
25     concentrations of 5, 29, and 33 ppb, diameter growth decreased by 29% in the highest O3
26     treatment, but height growth was unaffected (Bortier et al., 2000b). A 2-month study of hybrid
27     poplar (Populus tremuloides x P. tremula) in a free air exposure system in Finland with 7-h
28     mean O3 concentrations of 30 and  38 ppb found a 6% decrease in height with no effect on
29     biomass (Oksanen et al., 2001). Eastern cottonwood cuttings in pots buried in the ground with
30     drip irrigation were exposed to ambient O3  at several sites in and near New York City in three
31     2-month experiments during three summers (Gregg et al., 2003). Ozone concentrations were

       January 2005                             9-236       DRAFT-DO NOT QUOTE  OR CITE

-------
 1     lower at urban sites than at rural sites within 100 km of the urban sites.  Total biomass growth
 2     was greater in urban than rural sites, with a strong linear decrease in biomass with increasing O3
 3     across all sites and years (r2 = 93).  Total biomass decreased 33% with 12-h mean O3 levels of
 4     38 ppb compared to 23 ppb.  Multiple regression analysis showed no significant  temperature
 5     effect on biomass; therefore, the authors suggested that O3 exposures was the most likely
 6     explanation for the reduced biomass in rural areas.  The overall growth reductions and the
 7     variation among genotypes seen on  all of the above aspen and poplar studies is similar to those
 8     previously reported in three OTCs studies summarized in the previous criteria document
 9     (Table 9-26 of U.S. Environmental Protection Agency, 1996).
10          Black cherry seedlings grown in pots were exposed in OTCs in the Great Smoky Mountain
11     National Park in Tennessee to O3 treatments ranging from CF to 2x-ambient in two experiments
12     during 1989 and 1992 (Neufeld et al., 1995).  Ozone exposure, expressed as SUM06, ranged
13     from 0 to 40.6 ppm-h in 1989, and from 0 to 53.7 ppm-h in 1992. Corresponding AOT40 values
14     were 0.0 to 28.3 ppm-h in 1989, and 0 to 40.4 ppm-h in 1992. In 1989, total biomass was
15     decreased in the 1.5x-ambient treatment by 18% and in the 2x-ambient treatment by 38%.
16     In 1992, total biomass was decreased in the 1.5 x-ambient treatment by 27%, and in the
17     2x-ambient treatment by 59%. In this study, SUM06 and AOT40 provided better fits than did
18     SUMOO with Weibull regression analyses to log-transformed biomass data.  Although a Weibull
19     model was used, responses to O3 expressed as SUM06 and AOT40 appeared to be linear. The O3
20     exposures in the 1.5 x-ambient and 2 x-ambient treatments were reported to be similar to those
21     for a site near Charlotte, NC in a high-O3 year (1988).  In a 2-year experiment in  OTCs in Ohio,
22     seedlings of black cherry,  sugar maple, and yellow poplar were exposed to O3 treatments with
23     SUMOO values ranging from 16 to 107 ppm-h in 1990 and 31 to 197 ppm-h in 1991 (Rebbeck,
24     1996). After two seasons  of exposure, only black cherry showed a growth decrease: total
25     biomass was reduced by 32% in the 2x-ambient O3 treatment compared to the CF treatment; root
26     biomass was decreased by 39%. These results contrast with those of a previous study with black
27     cherry seedlings in which  significant biomass reductions with exposures up to 2 x-ambient were
28     not observed (7-h mean = 21 to 97 ppb), perhaps because of the small sample size (3 seedlings
29     per chamber (Samuelson,  1994) in the earlier study.
30          A multiyear study of effects  of O3 on both seedling and mature (30-year-old) red oak trees
31     was conducted in Norris, TN in large OTCs with three replicates per O3 treatment.  Trees were

       January 2005                             9-237       DRAFT-DO NOT QUOTE OR CITE

-------
 1     exposed for 3 years to CF, 1 x-ambient, and 2x -ambient treatments, with the following
 2     O3 exposures:  SUM06 for 3 years = 0, 29, 326 ppm-h; SUMOO for 3 years = 147, 255, and
 3     507 ppm-h. The net photosynthetic rate in mature trees was reduced by 25% in the ambient
 4     treatment and by 50% in the 2x-ambient treatment (Hanson et al., 1994; Samuelson and
 5     Edwards, 1993; Wullschleger et al., 1996). Despite these large decreases, no significant effects
 6     on stem increment at the base, stem increment in the canopy,  or leaf mass were observed for the
 7     mature trees (Samuelson et al., 1996).  Similarly, seedling biomass was not significantly reduced
 8     by O3 exposure. The difficulty in replicating experiments with mature trees makes it difficult to
 9     detect changes in growth or biomass.  However, the mean values of the stem increment at the
10     base and within the canopy in the ambient treatment were larger than those in the CF treatment,
11     although those in the 2x-ambient treatment were lower. Therefore this study of mature trees
12     does not provide evidence that these ambient concentrations reduced aboveground tree growth,
13     even after 4-years exposure.
14          Sugar maple seedlings were exposed for 3 years to ambient, 1.7x-ambient, and 3 x-ambient
15     O3 treatments at both high-light (3 5% of ambient) and low-light levels (15% of ambient) (Topa
16     etal., 2001). This experiment was conducted in large OTCs near Ithaca, NY.  Over the 3 years,
17     O3 exposures expressed as SUMOO for the three treatments were 88, 126, and 185 ppm-h. After
18     3 years, total seedling biomass in the 3 x-ambient treatment was reduced by 64% and 41% in the
19     low- and high-light treatments, respectively (compared to the 1 x-ambient treatment). The larger
20     reduction of biomass under low-light conditions suggests that seedlings growing under closed
21     canopies may be substantially more sensitive to O3 than are seedlings exposed to higher-light
22     levels in gaps or clearings.  These results differ from other studies in which seedling biomass
23     was unaffected by exposure to SUMOO values of 304 ppm-h over 2 years (Rebbeck, 1996) or
24     591 ppm-h  over 3  years (daytime mean of 40.7) (Laurence et  al., 1996). However the latter two
25     studies used much higher light levels, which may have reduced the O3 effect, based on the results
26     of Topa etal. (2001).
27          Although most studies demonstrate that O3 decreases biomass growth, occasional results
28     indicate that O3 can increase growth of some portions of woody perennials.  When Casselman
29     plum trees near Fresno, CA were exposed to O3 in large, rectangular OTCs to three O3 treatments
30     (CH, 1 x-ambient, and an above-ambient O3 treatment) for 4 years (12-h mean = 31, 48, 91
31     ppm-h), stem increment increased 14% in the highest O3 treatment compared to the CH

       January 2005                            9-23 8       DRAFT-DO NOT QUOTE OR CITE

-------
 1     treatment; and this effect was statistically significant (Retzlaff et al., 1997).  However, fruit yield
 2     decreased in this treatment by 42% and also decreased by 16% in the 1 x-ambient-O3 treatment.
 3     Root growth was not measured in this study. Hence, the increase in stem diameter may have
 4     been at the expense of other organs. However, in a fifth year, all plants were exposed to
 5     1 x-ambient O3, and there were no differences in fruit yield, suggesting that trees were able to
 6     recover to some extent from the effects of O3 exposure in prior years.
 7          When yellow poplar seedlings were exposed to O3 concentrations up to SUMOO values of
 8     107 ppm-h in one year and 197 ppm-h in a second year, no effects on biomass were observed
 9     (Rebbeck, 1996).
10          Many studies have demonstrated that root growth is more sensitive to O3 exposure than is
11     stem growth. For example, in a study with black cherry seedlings exposed in OTCs in
12     Tennessee in 1989, root biomass in a 2 x-ambient treatment was decreased by 42%, while stem
13     biomass was decreased by only 24%. However, in a second experiment in 1992, root and stem
14     growth reductions in the 2x-ambient treatment were similar (65% versus 60%) (Neufeld et al.,
15     1995). In Finland, reduced root growth was found for a number of clones of silver birch
16     (Oksanen and Saleem, 1999). After 5 years, root growth was decreased by 33%, but shoot
17     growth was not affected by O3 exposures of a 7-h mean of 15 ppm-h over 5 years in a FACE
18     system (Oksanen, 2001). When first-year poplar seedlings (P. tremuloides) were exposed in
19     OTCs to two O3 concentrations and six N concentrations, the root/shoot ratio was decreased soon
20     after exposure to O3 in all N treatments, even though O3 effects on total biomass were not
21     detected in the low-N and very high-N treatments (Pell  et al., 1995). These results suggest that
22     effects on the root/shoot ratio occur before significant growth effects arise. In a series of OTC
23     experiments lasting 1 to 3 years at 3 different elevations in Switzerland, fine root growth in
24     European beech was found to be more sensitive to O3 than was shoot or total biomass (Braun and
25     Fluckiger, 1995). Although the estimated effect of O3 on fine root biomass was similar to that
26     for total biomass, fine root biomass was significantly decreased at AOT40 (24-h) values of only
27     3 ppm-h, while total biomass was not significantly decreased until AOT40 values reached 30 to
28     40 ppm-h.
29
       January 2005                             9-239       DRAFT-DO NOT QUOTE OR CITE

-------
 1      9.6.5.3  European Critical Levels
 2           In Europe, a Level I critical level has been set for forest trees based on OTC studies of
 3      saplings.  This level is discussed here because it was based on a deciduous tree species.
 4      For consistency with the approach used for crops, an AOT40 index value was selected. A few
 5      studies have shown that O3 can be taken up by tree species at nighttime, e.g., young birch trees
 6      (Matyssek et al., 1995).  However, because most evidence suggests that O3 deposition at
 7      nighttime is low (Coe et al., 1995; Rondon et al., 1993), a value for only daylight hours was
 8      selected in Europe (Fuhrer et al., 1997; Karenlampi and Skarby, 1996). European beech was
 9      selected for development of a Level I critical level, because data from several studies were
10      available for this species and because deciduous tree species were judged to be more sensitive to
11      O3 compared to evergreen tree species (Fuhrer et al., 1997; Karenlampi and Skarby, 1996).
12      A critical level  was defined as an AOT40 value of 10 ppm-h for daylight hours for a 6-month
13      growing season (Karenlampi and Skarby, 1996). However, other studies have shown that other
14      species such as silver birch may be more sensitive to O3 than beech (Paakkonen et al., 1996).  As
15      discussed for annual plants above (Section 9.6.4.6), Level I critical values are not designed for
16      making quantitative estimates of the O3 effects on vegetation at the regional scale. For long-
17      lived perennials, additional problems complicate extrapolation. As discussed below (Section
18      9.6.5.7), considerable scaling is required to extrapolate from experiments conducted with tree
19      seedlings to estimate effects on mature trees in forests. Because of these scaling issues, there  is
20      greater uncertainty in estimating effects on forest trees than on annual plants such as crops.
21      While some information is available for addressing issues such as scaling from seedlings to
22      mature trees and estimating O3 uptake, this information may still be insufficient for developing a
23      Level II approach that can provide quantitative estimates of forest growth losses due to O3
24      exposure (Broadmeadow, 1998).
25
26      9.6.5.4  Summary of Effects on Deciduous Woody Species
27           Recent evidence from free air exposure systems and OTCs supports results observed
28      previously in OTC studies (Table 9-16, Figure 9-1). Specifically, a series of studies undertaken
29      using free air O3 enrichment in Rhinelander, WI (Isebrands et al., 2000, 2001) showed that
30      O3-symptom expression was generally similar in OTCs, FACE, and ambient-O3 gradient sites,
31      supporting the previously observed variation among aspen clones using OTCs (Karnosky et al.,

        January 2005                              9-240       DRAFT-DO NOT QUOTE OR CITE

-------
 1      1999). New evidence is also available comparing various aspects of O3 sensitivity between
 2      seedlings and mature trees of some species, notably red oak. As has been observed in previous
 3      criteria documents, root growth is often found to be the most sensitive indicator in terms of
 4      biomass response to O3.
 5           Results since 1996 support the conclusions of the 1996 AQCD (U.S. Environmental
 6      Protection Agency, 1996) that individual deciduous trees are generally less sensitive to O3 than
 7      are most annual plants, with the exception of a few genera such as Populus, which are highly
 8      sensitive. However, the data presented in Table 9-19 suggest that ambient exposures that occur
 9      in different regions of the United States can sometimes reduce the growth of seedlings of
10      deciduous species. Results from multiple-year studies  sometimes show a pattern of increasing
11      effects in subsequent years. Although, in some cases, growth decreases due to O3 become less
12      significant or even disappear over time.  While some mature trees show greater O3 sensitivity in
13      physiological parameters such as net photosynthetic rate compared to seedlings, these effects
14      may not translate into measurable reductions in biomass growth. Because even multiple-year
15      experiments do not expose trees to O3 for more than a small fraction of their life span and
16      because competition may, in some cases, exacerbate the effects of O3 on individual species,
17      determining the effects on mature trees remains a significant challenge. Effects on mature trees
18      under natural conditions are discussed after the review of evergreen species below and more
19      fully in Section 9.7, in the context of extrapolating from controlled studies to forest ecosystems.
20
21      9.6.5.5   Evergreen Woody Species
22           Most investigations have shown evergreen tree species to be less sensitive to O3 compared
23      to deciduous species (U.S. Environmental Protection Agency, 1996).  For example, exposure-
24      response functions based on 23 experimental cases of seedling response to O3, suggest that a
25      SUM06 exposure for 3 months of 42.6 ppm-h would protect evergreen species from a 10%
26      growth loss in 50% of the cases (Table 9-16). For deciduous species, the corresponding  SUM06
27      value was 31.5 ppm-h (Table 9-16). As another example, experiments in the Great Smoky
28      Mountains National Park found black cherry seedlings to demonstrate substantial  decreases in
29      biomass, as discussed above and shown in Table 9-19 (Neufeld et al., 1995). However, exposure
30      for up to three growing seasons did not decrease the biomass of eastern hemlock, Table
        January 2005                             9-241        DRAFT-DO NOT QUOTE OR CITE

-------
 1      Mountain pine, and Virginia pine seedlings exposed to O3 under similar conditions in this
 2      location, as shown in Table 9-19 (Neufeld et al., 2000).
 3           As for deciduous species, there is a substantial range in sensitivity among evergreen
 4      species.  As discussed above for deciduous species, a risk analysis was undertaken to predict tree
 5      biomass growth reductions due to O3 based on exposure-response equations for tree seedlings
 6      combined with species distribution across the eastern United States and interpolated O3 exposure
 7      (Hogsett et al., 1997). While some species such as Virginia pine were predicted to be affected
 8      only slightly even in a high O3 year, the growth of sensitive evergreen species such white pine
 9      was predicted to be reduced by 5% in a lower-than-average O3 year and 10% in a high ozone
10      year across 50% of its range (Andersen et al., 1997).  The remainder of this section discusses
11      experimental results for evergreen species in the order shown in Table 9-20.
12           Douglas fir seedlings were exposed to elevated O3 concentrations in a free air zonal air
13      pollution system in British Columbia, Canada for two growing seasons with 12-h mean values in
14      1988 of 18 to 41 ppb and in 1989 of 27 to 66 ppb (Runeckles and Wright, 1996). Although
15      substantial variation was seen in effects among the different treatments, there was a significant
16      decrease in the growth of the second flush weight  in the second year, with reductions of 55% at
17      the highest O3 exposure, based on a linear regression. This result contrasts with the lack of
18      effect seen in a previous study with seedlings of this species grown in pots for 134 days and
19      exposed to 7-h mean O3 concentrations up to 71 ppb (Table 9-30 in U.S. Environmental
20      Protection Agency, 1996).
21           First-year loblolly pine seedlings of 53 open-pollinated families were exposed to
22      1 x-ambient O3 in OTCs for a single growing season, and average growth volume was decreased
23      by 14% compared to a CF treatment (McLaughlin et al., 1994). The 1 x-ambient O3 exposure in
24      this study expressed as 24-h SUMOO was 137 ppm-h, and the CF treatment reduced O3 by 47%.
25      In this study, the root/shoot ratio was decreased significantly in six of the nine families
26      examined. Exposure to O3 with SUM06 values up to 162 ppm-h and  132 ppm-h in 2 successive
27      years in OTCs had no effect on seedlings grown in competition with various species of grasses
28      and forbs (Barbo et al., 2002). Exposure of 3-year-old seedlings to O3 exposures of up to
29      2.5x-ambient (12-h mean of 98 ppb) also had no significant effect (Anttonen et al., 1996).
30      Information summarized in the 1996 AQCD (U.S. Environmental Protection Agency, 1996),
31      indicated that significant effects on seedling growth were observed in several studies of

        January 2005                             9-242        DRAFT-DO NOT QUOTE OR CITE

-------
 1      seedlings exposed to elevated O3 concentrations for one or more years. Several studies,
 2      including that of McLaughlin et al. (1994), demonstrate considerable variation in O3 sensitivity
 3      among different genotypes of loblolly pine.
 4           For ponderosa pine seedlings, the  1996 AQCD reviewed a number of studies with
 5      exposures to elevated O3 concentrations for one to three growing seasons (U.S. Environmental
 6      Protection Agency, 1996).  Recent similar studies support the earlier results (Table 9-20);
 7      (Andersen et al., 2001; Takemoto et al.,  1997). The  1996 criteria document also discussed at
 8      some length the ongoing work examining effects of O3 across a naturally-occurring O3 gradient
 9      in the San Bernardino Mountains in California.  Since that time, much research on ponderosa
10      pine has focused on interactive effects of additional stresses such as nitrogen, and effects of O3
11      on physiological parameters (Sections 9.4, 9.7). Effort has also focused on the effects of O3 on
12      root growth because such effects could alter sensitivity to drought or nutrient stress. Ecosystem
13      level effects of ozone are discussed further in Section 9.7, but some information relevant to
14      exposure-response relationships is discussed below.
15           For several tree species, O3 has been shown in  experimental studies with  seedlings to
16      reduce root growth more than shoot growth (U.S. Environmental Protection Agency, 1996).
17      Ponderosa pine has been shown to be sensitive to O3, and studies with seedlings have shown
18      reduced root growth, decreases root/shoot ratios, and decreased allocation to roots (Andersen
19      et al.,  1991; 1997; Andersen and Rygiewicz, 1995; Andersen and Scagel, 1997; Tingey et al.,
20      1976). More recently, data from a long-studied pollution gradient in the San Bernardino
21      Mountains of southern California suggests that O3 substantially reduces root growth in natural
22      stands of ponderosa pine.  Root growth in mature trees was decreased at least 87% in a high
23      pollution site as compared to a low pollution site (Grulke et al., 1998), and a similar pattern was
24      found in a separate study with whole tree harvest along this gradient (Grulke and Balduman,
25      1999). Because other potential influences on root growth, including shading by competing trees,
26      soil temperature, soil moisture, phenology, were not correlated with the observed pattern of
27      reduced root growth, the authors conclude that O3 was the cause of the observed decline in root
28      growth.  Further results of field investigations with ponderosa pine and other pine species native
29      to California are discussed below under the heading  "Scaling experimental data on long-lived
30      species to field conditions" as well as in Section 9.7.
        January 2005                             9-243        DRAFT-DO NOT QUOTE OR CITE

-------
 1           Table Mountain pine, Virginia pine, and eastern hemlock seedlings were exposed to
 2      various levels of O3 (CF to 2 x ambient) in OTCs for in a series of experiments two or three
 3      years in Great Smoky Mountains National Park in Tennessee (Neufeld et al., 2000).  There were
 4      no statistically significant effects of O3 exposure on stem or root biomass, and only slight effects
 5      on the biomass of the oldest needles in Table Mountain pine in the 2* ambient treatment.
 6           As reviewed in the 1996 criteria document, studies of the response of red spruce to
 7      O3 exposures generally have not found effects on growth of seedlings or saplings, even after
 8      exposure to high concentrations (12-h mean of 90 ppb) for up to 4 years. A more recent report
 9      confirms that this slow-growing species is O3 insensitive for at least several years (Laurence
10      etal., 1997).
11           For perennial vegetation, cumulative effects over more than one growing season may be
12      important.  For three-year-old Norway spruce in Sweden, exposure to elevated O3 for three
13      growing seasons decreased total biomass by 18% and stem biomass by 28% (Karlsson et al.,
14      1995).  However, after a fourth season of exposure, total biomass decreased significantly by only
15      8% (Karlsson et al., 2002). In this experiment, the O3 exposures expressed as 12-h mean values
16      averaged over four growing seasons were 12 and 44 ppb for the CF and 1.5 x-ambient treatments,
17      respectively; and AOT40 values were 2 and 23 ppm-h, respectively.  Despite 4 years of
18      exposure, this experiment did not demonstrate a consistent trend in the O3 effect on biomass to
19      indicate a significant carry-over effect.
20
21      9.6.5.6  Summary of Effects on Evergreen Woody Species
22           In summary, the O3 sensitivity of different genotypes within species and between species
23      of evergreen vegetation varies widely.  Based on studies with seedlings in OTCs, major species
24      in the United States are generally less sensitive than are most deciduous trees, and slower-
25      growing species are less sensitive than faster-growing ones. However, exposure to ambient O3
26      may reduce the growth of seedlings of commonly occurring species.  Because tree species are
27      long-lived, most experiments have only covered a very small portion of the life span of a tree,
28      making estimating of any effect on mature trees difficult. Considerations for scaling the results
29      of seedling studies to mature forest trees as well as additional information from field surveys and
30      studies of mature trees under natural conditions are discussed below and in Section 9.7.
31

        January 2005                              9-244       DRAFT-DO NOT QUOTE OR CITE

-------
 1      9.6.5.7  Scaling Experimental Data to Mature Trees
 2           As compared with annual crop species, it is much more difficult to define appropriate
 3      exposure-response relationships for tree species.  For annual species, an experiment may cover
 4      the whole life span of the plant, but it is not feasible to provide controlled-exposure conditions
 5      for long-lived plants for any significant portion of their life spans. Most studies have used small
 6      seedlings because they are manageable under experimental conditions; but seedlings and mature
 7      trees may have different sensitivities to O3.  For perennial species, effects of O3 may accumulate
 8      over more than 1 year, and may interact with other stresses such as drought stress over multiple
 9      growing seasons.  As for annual species (Section 9.4.2), substantial variability occurs among
10      evergreen genotypes and this variation may interact with other stress responses differently in
11      different landscapes and regions.  Despite these difficulties, investigators have addressed some
12      of these issues since the 1996 AQCD (U.S. Environmental Protection Agency, 1996). New
13      information is available on the response of mature evergreen trees to O3 under field conditions,
14      and models based on tree physiology and stand dynamics have been used to predict O3 effects on
15      forest stands and regions.  The following issues are reviewed briefly below: (1) interaction of
16      drought and O3 stress, (2) scaling  data from seedlings to mature-tree studies. Two additional
17      scaling issues are addressed in Section 9.7:  (1) scaling data to forest stands, and (2) scaling data
18      to ecosystems and regions.
19
20      9.6.5.7.1 Interactive Effects of Drought and Ozone
21           Many interacting factors may influence the effect of O3 on vegetation. For crop plants,
22      environmental conditions are often managed such that nutrients and water are not strongly
23      limiting; but for native vegetation, including most perennial species, such factors are likely to
24      limit growth.  The effects of interacting stresses on vegetation are reviewed in Section 9.4.
25      However, because drought is common in many forests, and because there is substantial evidence
26      that it alters the response of trees to O3, it is discussed in this section in the context of
27      determining exposure-response relationships for trees.
28           Controlled experiments with seedlings provide direct evidence that drought can reduce the
29      impact of O3. For example, for 3-year-old Norway  spruce in  Sweden, exposure to elevated
30      O3 for three growing seasons decreased total biomass by 18% and stem biomass by 28%
31      (Karlsson et al., 1995). However, for draughted trees, both total and stem biomass decreased

        January 2005                             9-245        DRAFT-DO NOT QUOTE OR CITE

-------
 1      only 5%, with a statistically significant interaction with O3 for stem biomass.  Yet after a fourth
 2      season of exposure, there was no longer any interaction between drought and O3, while there was
 3      a significant decrease of 8% in the biomass when both drought and well-watered data were
 4      combined (Karlsson et al., 2002). In this study, seedlings were grown in sand in 120 L pots and
 5      for the drought treatment, water was withheld for 4 weeks during the first year and for 7 to
 6      8 weeks during each of the last 3 years. In this experiment, the O3 exposures expressed as 12 h
 7      seasonal daylight mean averaged over four growing seasons were 12 and 44 ppb for the CF and
 8      1.5x-ambient treatments respectively.  Over this period, the AOT40 values for the treatments
 9      averaged 2 and 23 ppm-h respectively. Despite 4 years of exposures, this experiment did not
10      demonstrate a consistent trend in drought O3 interactions. The difference in effects seen between
11      the third and fourth season suggest that scaling drought-O3 interactions from seedlings to mature
12      trees may be difficult. However, evidence from biomonitoring surveys supports an interaction
13      between drought and O3 effects, at least for foliar symptoms. In systematic surveys of foliar
14      symptoms on species selected as biomonitors throughout much of the eastern United States,
15      symptoms were more common and more severe in areas with high O3 concentrations (Smith
16      et al., 2003). However,  in very dry years, such as 1999, the occurrence and severity of
17      symptoms was greatly reduced, even in areas with high ambient O3 concentrations.
18
19      9.6.5.7.2  Scaling from Seedlings to Mature Trees
20           Because most experiments are conducted with seedlings, various methods are required to
21      scale experimental data  on seedlings to mature trees. An overview of physiological differences
22      between young and old plants, and the consequences of these differences for O3 sensitivity, is
23      provided in Section 9.4.5.3. The discussion below focuses on information relevant to developing
24      exposure-response relationships for mature trees. Information from a few experimental studies,
25      as well as scaling efforts based on physiological characteristics incorporated into models, are
26      discussed in Section 9.7.
27           Although most studies continue to examine the effects of O3 on seedlings, during the 1990s
28      some studies examined the effects of O3 on the response of mature trees.  Studies of mature trees
29      demonstrate differences in some aspects of O3 sensitivity between seedlings and mature trees.
30      For some species, such as  red oak, seedlings are less sensitive to O3 than are mature trees
31      (Hanson et al., 1994; Samuelson and Edwards, 1993; Wullschleger et al., 1996). Both red oak

        January 2005                              9-246        DRAFT-DO NOT  QUOTE OR CITE

-------
 1      seedlings and genetically related mature trees were exposed to CF, 1 x -ambient, or 2 x-ambient
 2      O3 exposures in OTCs in Tennessee for two growing seasons (Hanson et al., 1994). Nine large
 3      chambers (4.6 x 8.2 m) were used to enclose individual mature trees and standard EPA-style
 4      OTCs were used for potted seedlings. Ozone exposures expressed as a 24-h SUMOO were 34,
 5      79, and 147 ppm-h in  1992 and 37, 95, 188 ppm-h in 1993 for the sub-ambient, and 2x-ambient
 6      treatments. Mature trees had a greater light-saturated net photo synthetic rate and stomatal
 7      conductance compared to seedling foliage at physiological maturity. By the end of the growing
 8      season, exposure to 1 x-ambient and 1 x-ambient O3 reduced the light-saturated net
 9      photosynthetic rate and stomatal conductance of mature trees by 25 and 50%, respectively,
10      compared with the CF treatment (35 ppm-h).  In seedlings, however, light-saturated net
11      photosynthetic rate and stomatal conductance were less affected by O3 exposure.  The authors
12      concluded that extrapolations of the results of seedling-exposure studies to foliar responses of
13      mature forests without considering differences in foliar anatomy and stomatal response between
14      juvenile and mature foliage may introduce large errors into projections of the O3 responses of
15      mature trees.
16           In a study of ponderosa pine in California, seedlings and branches of mature trees
17      (in branch chambers) were exposed to O3 concentrations of 0.5-, 1-, and 2x-ambient O3
18      concentrations (Momen et al., 1997).  Net photosynthetic rate of 1-year-old, but not current-year,
19      foliage was reduced in mature trees while not significantly reduced in seedlings.  This effect was
20      not due to alteration of stomatal conductance by O3.  This result contrasts with those with earlier
21      studies of red spruce (Rebbeck et al.,  1993).
22           In contrast to the findings for red oak and ponderosa pine, giant sequoia seedlings had
23      higher rates of stomatal conductance,  CO2-exchange rate, and dark respiration than did mature
24      trees (Grulke and Miller, 1994). As compared to older trees, stomatal conductance was more
25      than 7-fold greater in current-year and 4-fold greater in 2-year-old seedlings (Grulke and Miller,
26      1994). The authors concluded that giant sequoia seedlings are sensitive to atmospheric O3 until
27      ~5 years of age. Low conductance, high water use efficiency, and compact mesophyll all
28      contribute to a natural O3 tolerance, or to O3 defense, or to both, in the foliage of older trees.
29      Similarly, lower stomatal conductance was found in mature Norway spruce in Austria compared
30      to seedlings grown with optimal water and nutrients in a growth chamber (Wieser, 1997). In this
31      study, net photosynthetic rate was less sensitive to added O3 in mature trees compared to

        January 2005                              9-247       DRAFT-DO NOT QUOTE OR CITE

-------
 1      seedlings.  In a related study, the average rate of O3 uptake of 17-year-old trees
 2      -0.6 nmol nT2 s"1, decreasing linearly in older trees, such that rates were only ~0.1 nT2 s"1 in
 3      216-year-old trees (Wieser et al., 1999).
 4           Based on a review of studies of stomatal conductance in both seedlings and mature trees,
 5      Samuelson and Kelly (2001) concluded that O3 uptake in oak species, black cherry, sugar maple,
 6      and American beech averaged 47% lower in potted seedlings than in mature trees.  For evergreen
 7      species, they concluded that O3 uptake in seedlings averaged 26% higher than in mature trees.
 8      They also suggested that artifacts introduced by growth in pots confound these differences that
 9      exposure-response functions derived from seedlings grown in situ are more applicable to mature
10      trees than are studies of seedlings grown in pots (Samuelson and Kelly, 2001).
11           As discussed above for annual vegetation, it has long been noted that internal O3 dose is
12      more appropriate than external O3 exposure for assessing the effects of O3 on vegetation, because
13      effects occur primarily via the uptake of O3 through the stomata (Section 9.3.2).  However,
14      external O3 exposure sometimes has been shown to explain O3 effects as well or better than
15      calculated internal O3 dose.  For ponderosa pine, Grulke and others (2002) found little difference
16      in the response of net photosynthetic rate and stomatal conductance to O3 exposure as compared
17      to calculated O3 uptake; and estimated O3 uptake by ponderosa pine and O3 exposure at several
18      sites were highly correlated (r2 = 0.92). For red oak, Hanson and others (1994) found that
19      SUMOO explained 83% of the variance in the response of light-saturated photosynthetic rate to
20      O3 levels, while estimated internal dose explained only 76% of the variance. In this same study,
21      SUM06 explained only 49% of the variance.  Due to genetic variation or other factors, individual
22      mature trees will vary in their response to similar O3 exposures. For example, in 125-year-old
23      giant sequoia trees exposed to -230  ppm-h of O3 in branch chambers, O3 uptake in one
24      individual  was -5 mmol nT2, while in another it was -9.5 mmol nT2 (Grulke et al., 1996).
25           Based on these results, stomatal conductance, O3 uptake, and O3 effects cannot be assumed
26      to be equivalent in seedlings and mature trees.  In general, mature deciduous trees are likely to
27      be more sensitive to O3 compared to seedlings,  while mature evergreen trees are likely to be less
28      sensitive than seedlings. However, even when  differences in physiological traits occur,
29      concomitant  effects on stem growth  may not be detected in the field.  Additionally, complex
30      interactions may occur between environmental  conditions and O3 responses;and artifacts may
31      occur for seedling studies, especially for seedlings grown in pots.  Section 9.7 further discusses

        January 2005                             9-248       DRAFT-DO NOT QUOTE OR CITE

-------
 1      issues that must be addressed when scaling data from individual mature trees to forests and
 2      regions.
 3
 4      9.6.6  Studies with the Chemical EDU
 5           The chemical EDU (7V-[2-(2-oxo-l-imidagolidinyl)ethyl]-7V-phenylurea) has been used
 6      with the goal of protecting plants from O3 effects without controlling O3 exposure (Table 9-21)
 7      (U.S. Environmental Protection Agency, 1986, 1996). The use of EDU has the potential to be a
 8      low-cost,  practical method of evaluating ambient O3 exposures on plants grown under natural
 9      conditions without the limitations imposed by methodologies such as OTCs (Section 9.2). EDU
10      can be applied easily as a foliar spray or a soil drench and, with more effort,  can be injected
11      directly into plant stems.  However, because EDU is phytotoxic, and may have effects on plants
12      other than antioxidant protection, it is crucial that the correct dosage for protection from O3, be
13      determined without the direct effects of EDU. Unfortunately, although many studies with EDU
14      have been conducted in recent  decades, very few have used multiple EDU application levels
15      along with multiple O3 exposures to characterize the EDU system for a given plant species.
16      Therefore, the text of this section focuses on how data from existing studies can be used for
17      developing or validating exposure-response relationships, rather than reviewing results of all
18      individual studies. Data from individual studies on O3 exposure, EDU application rates, and the
19      effects  of EDU are presented in Table 9-21.  In addition to EDU, sodium erythorbate has been
20      used in a few studies as a protectant chemical. Since very few published studies have used
21      sodium erythorbate and attempts to establish appropriate doses for individual species  are even
22      more limited, the use of this chemical is not reviewed here.
23           The phytotoxicity of EDU is well known, and the point has been made repeatedly that for
24      a particular species or cultivar, tests under a range of environmental conditions and O3 exposures
25      must be made to establish the efficacy of EDU for quantifying O3 effects (Heggestad, 1988;
26      Kostka-Rick and Manning,  1992). A recent study has shown that even low concentrations of
27      EDU (8 to 32 mg I/1 soil), decreased bean yield under low O3 exposure (7-h mean of 19 ppb) in
28      CF OTCs (Miller et al., 1994).   This study also demonstrated that phytotoxicity  (both foliar
29      symptoms and growth effects)  can differ even in the same series of experiments, apparently due
30      to changes in environmental conditions, and that EDU can suppress yield at application rates that
31      do not always cause foliar symptoms (Miller et al., 1994).  Finally, this study found that EDU

        January 2005                              9-249        DRAFT-DO NOT QUOTE OR CITE

-------
 1      altered biomass partitioning by increasing vegetative growth and decreasing reproductive
 2      growth.  A study of bean grown in OTCs in Germany found that EDU treatment in CF OTCs
 3      significantly increased yield, while EDU had no significant effect on yield in other O3 treatments
 4      (Brunschon-Harti et al., 1995).  In this study, O3 significantly reduced the mass of pods, shoots,
 5      and roots. EDU increased root, leaf, and shoot mass across O3 treatments. However, there was
 6      only a significant interaction with O3 for root mass.  This study indicates that EDU can stimulate
 7      above-ground growth and/or delay senescence regardless of O3 treatment. Together, these
 8      studies suggest that EDU has effects other than its antioxidant protection and phytotoxicity.
 9           Recent studies have also shown that EDU does not always have greater effects at higher O3
10      exposures. For bean grown in pots in studies in Spain and the Netherlands, EDU increased pod
11      yield (Ribas and Penuelas, 2000; Tonneijck and Van Dijk, 1997). However, this effect was not
12      greater at sites with higher O3 exposure despite consistent experimental protocols at all sites,
13      including growing the same cultivar in pots with adequate water (Ribas and Penuelas, 2000;
14      Tonneijck and Van Dijk, 1997).  Such results suggest that it may be difficult to quantify
15      ambient O3 effects using EDU, because the amount of plant growth or yield expected at a low
16      (background)  O3 concentration cannot be inferred from EDU-treated plants grown at locations
17      with higher O3 exposures.
18           Potted white clover exposed to ambient O3 at 12 sites throughout Europe over
19      three growing seasons was evaluated in a meta-analysis (Ball et al., 1998).  A soil  drench of
20      100 mL EDU  was applied every 2 weeks for 3 months per growing season. Only very weak
21      evidence of a linear relationship of the ratio of biomass in control versus EDU treated plants
22      across all sites (r2 = 0.16) was seen.  However, an artificial neural network (ANN) model
23      including VPD, temperature, longitude, year, and altitude explained much more of the variance
24      (r2 = 0.79).  The authors suggested that the greater sensitivity at certain sites in Germany may
25      have been due to occurrence of other pollutants. This meta-analysis indicates that EDU effects
26      may be influenced substantially by environmental factors.
27           In summary, the EDU method for assessing the impact of ambient O3 exposures is
28      potentially useful, because it provides a separate line of evidence than other methods.  However,
29      as has been pointed out by numerous authors, the system must be carefully characterized for
30      each species under different environmental conditions and different O3 exposures.
31      Unfortunately, such characterization has so far been limited,  although substantial progress has

        January 2005                              9-250        DRAFT-DO NOT QUOTE OR CITE

-------
 1      been made for radish (Kostka-Rick et al., 1993; Kostka-Rick and Manning, 1992, 1993).  Effects
 2      of EDU can include phytotoxicity and alteration of biomass partitioning, along with differential
 3      effects on the same species in different years or locations. Additionally, the degree of O3
 4      protection afforded by a certain application of EDU is difficult to quantify.  Thus, it is difficult to
 5      use data from existing EDU studies to develop exposure-response relationships or to quantify the
 6      effects of ambient O3 exposure. Despite these limitations, the EDU studies reviewed in previous
 7      criteria documents (U.S. Environmental Protection Agency, 1986, 1996) and the more recent
 8      studies summarized in Table 9-21 provide another line of evidence that ambient O3 exposures
 9      occurring in many regions of the United States may be reducing the growth of crops and trees.
10
11      9.6.7   Summary
12          Recently published data support the conclusions of previous criteria documents that there
13      is strong evidence that ambient concentrations of O3 cause injury and damage to numerous
14      common and economically valuable plant species. For annual vegetation, the data summarized
15      in Table 9-17 show a range of growth and yield responses both within and among species.
16      Nearly all of these data were derived from studies in OTCs, with only two studies using open-air
17      systems in the UK (Ollerenshaw et al., 1999; Ollerenshaw and Lyons, 1999). It is difficult to
18      compare studies that report O3 exposure using different indices, such as AOT40, SUM06, or 7-h
19      or 12-h mean values. However, when such comparisons can be made, the results of recent
20      research confirm earlier results summarized in the 1996 AQCD (U.S. Environmental Protection
21      Agency, 1996). A summary of earlier literature concluded that  a 7-h, 3-month mean of 49 ppb
22      corresponding to a SUM06 exposure of 26 ppm-h would cause 10% loss in 50% of 49
23      experimental cases (Tingey et al., 1991).  Recent data summarized in Table 9-17 support this
24      conclusion, and more generally indicate that ambient O3 exposures can reduce the growth and
25      yield of annual species.  Some annual species such as soybean are more sensitive, and greater
26      losses may be expected (Table 9-17).  Thus the recent scientific literature supports the
27      conclusions of the 1996 AQCD (U.S.  Environmental Protection Agency, 1996) that ambient O3
28      concentrations are probably reducing  the yield of major crops in the United States.
29          Much research in Europe has used the AOT40  exposure statistic, and substantial effort has
30      gone into developing Level-1 critical  levels for vegetation using this index. Based on regression
31      analysis of 15 OTC studies of spring wheat including one study from the United States and

        January 2005                              9-251        DRAFT-DO NOT QUOTE OR CITE

-------
 1      14 from locations ranging from southern Sweden to Switzerland, an AOT40 value of 5.7 ppm-h
 2      was found to correspond to a 10% yield loss, and a value of 2.8 ppm-h corresponded to a 5%
 3      yield loss (Fuhrer et al., 1997). Because a 4 to 5% decrease could be detected with a confidence
 4      level of 99%, a critical level of an AOT40 value of 3 ppm-h was selected in 1996 (Karenlampi
 5      and Skarby, 1996).
 6           In addition to likely reductions in crop yield, O3 may also reduce the quality or nutritive
 7      value of annual species. Many studies have shown effects of O3 on various measures of plant
 8      organs that affect quality, with most studies focusing on characteristics important for food or
 9      fodder. These studies indicate that there may be economically important effects of ambient O3
10      on the quality of crop and forage species.  Previous criteria documents have concluded that
11      visible symptoms on marketable portions of crops and ornamental plants can occur with seasonal
12      7-h mean O3 exposures of 40 to 100 ppb (U.S. Environmental Protection Agency, 1978, 1986,
13      1996).  The recent scientific literature does not refute this conclusion.
14           The use of OTCs may reverse the usual vertical gradient in O3 that occurs within a few
15      meters above the ground surface (Section 9.2). This reversal  suggests that OTC studies may
16      somewhat overestimate the effects of an O3 concentration measured several meters above the
17      ground. However such considerations do not invalidate the conclusion of the 1996 AQCD (U.S.
18      Environmental Protection Agency, 1996) that ambient O3 exposures (Tables  9-14 and
19      Table 9-22) are sufficient to reduce the yield of major crops in the United States.
20           As for single-season agricultural crops, yields of multiple-year forage crops are reduced at
21      O3 exposures that occur over large areas of the United States.  This result is similar to that
22      reported in the 1996 AQCD (U.S. Environmental Protection Agency, 1996).  When species are
23      grown in mixtures, O3 exposure can increase the growth of O3- tolerant species and exacerbate
24      the growth decrease of O3-sensitive species (e.g., Ashmore and Ainsworth, 1995; Fuhrer et al.,
25      1994).  Because of this competitive interaction, the total growth of the mixed-species community
26      may not be affected by O3 exposure (Ashmore and Ainsworth, 1995; Barbo et al., 1998; Fuhrer
27      et al., 1994). However, in some cases, mixtures of grasses and clover species have shown
28      significant decreases in total biomass growth in response to O3 exposure in studies in the United
29      States (Heagle et al.,  1989; Kohut et al., 1988) and in Sweden (Pleijel et al., 1996).  In Europe,
30      a provisional critical  level for herbaceous perennials of an AOT40 value of 7 ppm-h over
31      6 months has been proposed to protect sensitive plant species from adverse effects of O3.

        January 2005                             9-252       DRAFT-DO NOT QUOTE OR CITE

-------
 1           For deciduous tree species, recent evidence from free air exposure systems and OTCs
 2      supports results observed previously in OTC studies. For example, a series of studies
 3      undertaken using free air O3 enrichment in Rhinelander, WI (Isebrands et al., 2000, 2001)
 4      demonstated that O3-symptom expression was generally similar in OTCs, FACE, and also sites
 5      along an ambient O3 gradient, supporting the previously observed variation among aspen clones
 6      using OTCs (Karnosky et al.,  1999). As has been observed in previous criteria documents, root
 7      growth often is found to be the most sensitive biomass response indicator to O3.
 8           Results reported since 1996 support the conclusion of the 1996 AQCD (U.S.
 9      Environmental Protection Agency,  1996) that deciduous trees are generally less sensitive to
10      O3 than are most annual plants, with the exception of a few very sensitive genera such as
11      Populus and sensitive species such  as black cherry.  However, the data presented in Table 9-19
12      suggest that ambient O3 exposures that occur in the United States can potentially reduce the
13      growth of seedlings of deciduous species. Results from multiple-year studies  sometimes show a
14      pattern of increased effects in subsequent years. In some cases, however, growth decreases due
15      to O3 may become less significant or even disappear over time. While some mature trees show
16      greater O3 sensitivity in physiological parameters  such as net photosynthetic rate than do
17      seedlings, these effects may not translate into measurable reductions in biomass growth.
18      However, because even multiple-year experiments do not expose trees to O3 for more than a
19      small fraction of their life span, and because competition may in some cases exacerbate the
20      effects of O3 on individual species,  determining effects  on mature trees remains a significant
21      challenge.
22           In Europe, a Level I critical level has been set  for forest trees based on OTC studies of
23      European beech seedlings.  A critical level was defined as an AOT40 value of 10 ppm-h for
24      daylight hours for a 6-month growing season (Karenlampi and  Skarby, 1996). However, other
25      studies show that some species such as silver birch may be more sensitive to O3 compared to
26      beech (Paakkonen et al., 1996).
27           For evergreen tree species, as for other tree species, the O3 sensitivity of different
28      genotypes and different species varies  widely. Based on studies with seedlings in OTCs, major
29      species in the United States are generally less sensitive  than are most deciduous trees, and
30      slower- growing species are less sensitive than are faster-growing species. Interacting stresses
31      such as competition stress may increase the sensitivity of trees to O3. As for deciduous species,

        January 2005                              9-253        DRAFT-DO NOT QUOTE OR CITE

-------
 1      most experiments with evergreen species have only covered a small portion of the life span of a
 2      tree and have been conducted with seedlings, making estimating effects on mature trees difficult.
 3          For all types of perennial vegetation, cumulative effects over more than one growing
 4      season may be important, furthermore, studies for only a single season may underestimate
 5      effects. Mature trees may be more or less sensitive to O3 than are seedlings, depending on the
 6      species, but information on physiological traits may be used to predict some such differences.
 7      In some cases, mature trees may be more sensitive to O3 than seedlings due to differences in
 8      their gas exchange rates, growth rates, greater cumulative exposures, or due to the interaction of
 9      O3 stress with other stresses.
10
11
12      9.7  EFFECTS OF OZONE EXPOSURE ON NATURAL  ECOSYSTEMS
13      9.7.1   Introduction
14          The preceding section on species-level responses (9.6) provides a lead-in to address the
15      response of ecosystems to ozone (O3). The conclusion of the 1996 O3 AQCD was that aside
16      from the results from the San Bernardino Forest, there was no direct evidence  that O3 is altering
17      natural  ecosystems in the United States. This conclusion is generally valid today, except that our
18      understanding of the effects of O3 in the San Bernardino forest has been tempered by additional
19      understanding of the complicating role that N deposition plays in this system.  Despite the lack
20      of any new, direct information  linking O3 with ecosystem changes, numerous publications since
21      1996 have highlighted ways in which O3 may affect ecosystem structure and/or function.  This
22      section addresses new and (where appropriate) older literature in order to illustrate possible
23      shifts in energy or material flow through ecosystems as a result of O3 exposure.
24          An ecosystem is defined as comprising all of the organisms in a given area interacting with
25      the physical environment, so that a flow of energy leads to a clearly defined trophic structure,
26      biotic diversity, and cycling of materials between living and nonliving parts (Odum, 1963).
27      Individuals within a species and populations of species are the building blocks from which
28      communities and ecosystems are constructed.  Classes of natural ecosystems are distinguished
29      by their dominant vegetation form, e.g., tundra, wetland, deciduous forest and conifer forest.
30      Boundaries of ecosystems are delineated when an integral unit is formed by the physical and
        January 2005                             9-254        DRAFT-DO NOT QUOTE OR CITE

-------
 1      biological parts. Defined pathways for material transport and cycling, and for the flow of energy
 2      are contained within the integrated unit.
 3           Each level of organization within an ecosystem has functional and structural
 4      characteristics. At the ecosystem level, functional characteristics include but are not limited to
 5      energy flow; nutrient, hydrologic, and biogeochemical cycling; and maintenance of food chains.
 6      The sum of the functions carried out by ecosystem components provides many benefits to
 7      mankind as in the case of forest ecosystems (Smith, 1992).  These include food, fiber production,
 8      aesthetics, genetic diversity, and energy exchange.
 9           Ecosystems are functionally highly integrated.  Thus, changes in one part of an ecosystem,
10      such as the primary producer component, may have consequences for connected parts, such as
11      the consumer and decomposer components. For example, when needles are shed prematurely as
12      a result of O3 exposure, successional development of phyllosphere fungi inhabiting the surface of
13      ponderosa pine needles may be truncated (Bruhn, 1980). In addition, decomposer populations in
14      the litter layer may be capable of higher rates of decomposition, because the younger age classes
15      of needles falling from O3-damaged pines have higher nitrogen (N) content (Fenn and Dunn,
16      1989). Because ecological systems integrate the effects of many influences, the resulting effect
17      of O3 exposure may depend on co-occuring influences that predispose an ecosystem to stress
18      (Colls and Unsworth, 1992). One important change in our thinking since the 1996 O3 AQCD is
19      that at high levels of O3 exposure that are known to result in detectable plant responses
20      (> 250 ppm h accumulated over a growing season), N deposition must also  be considered as a
21      concurrent stressor.  Changes in N cycling and compartmentalization in the ecosystem result
22      from both O3 exposure and increased N deposition.
23           The vast majority of O3 effects literature addresses individual species responses (see
24      Section 9.7.4.3), which was also true in 1996 (U.S. Environmental Protection Agency, 1996).
25      This section  differs from the preceding one in that physiological stress of individual species is
26      considered only within the context of the natural ecosystem. Changes in function at the level of
27      the individual are propagated through higher levels of organization, resulting in changes in
28      ecosystem structure and function. However, since ecosystem-level responses result from the
29      interaction of organisms with one another and with their physical environment, it takes longer
30      for a change to develop to a level of prominence that can be identified and measured.  The
31      paucity of scientific literature on O3 effects at the ecosystem level is a result of both the

        January 2005                             9-255       DRAFT-DO NOT QUOTE OR CITE

-------
 1      complexity of ecological systems, and long response times. In addition, "indirect" effects of O3
 2      on plants (e.g., affecting the plants' ability to integrate environmental stresses) may be more
 3      important than direct effects on photosynthesis and respiration at the leaf level (Johnson and
 4      Taylor, 1988).
 5           A conceptual framework (see Table 9-22) suitable for organizing discussion of the effect of
 6      O3 on ecosystems was developed by the EPA Science Advisory Board (Young and Sanzone,
 7      2002). Their six essential ecological attributes (EEAs) include landscape condition, biotic
 8      condition, chemical and physical characteristics, ecological processes, hydrological and
 9      geomorphological processes, and natural disturbance regimes (listed as subsection headings
10      below and in Table 9-22). The major ecological effects, and gaps in our knowledge of O3
11      exposure effects at the ecosystem level are summarized at the end of this chapter.  While the
12      main focus is O3 effects newly described since the last Ozone Criteria Document (EPA, 1996),
13      many key historical papers are cited to demonstrate ecosystem response, particularly where they
14      remain the only examples in the literature. Although the vast majority of published studies focus
15      on individuals, five field examples and one FACE experiment have measured several ecosystem
16      components simultaneously to better understand ecosystem response to O3. We provide an
17      overview of these six studies up-front because they provide a context for the subsequent
18      discussion on possible ecosystem effects.
19
20      9.7.2 Case Studies
21      9.7.2.1  Valley of Mexico
22           The first evidence of air pollution impacts on vegetation in the Valley of Mexico (Mexico
23      City Air Basin) were observations of foliar injury symptoms in bioindicator plants attributed to
24      O3, PAN, SO2 and possibly other pollutants (de Bauer,  1972). Subsequently, O3 injury to foliage
25      and crowns of pine trees were reported in forests to the south and southwest of Mexico  City
26      (Krupa and de Bauer, 1976; de Bauer and Hernandez-Tejeda, 1986). Ozone is considered to be
27      the pollutant with the most severe impacts on vegetation within the Mexico City urban zone and
28      in forests downwind of the city. P. hartwegii is the most O3-sensitive pine species and is
29      severely impacted by high O3 exposures encountered to the south/southwest of the metropolitan
30      area (Miller et al., 2002).  The potential for O3 injury is particularly high in this area, because O3
31      levels are high during the summer rainy season when soil moisture  availability and stomatal

        January 2005                              9-256       DRAFT-DO NOT QUOTE OR CITE

-------
    Table 9-22. Essential Ecological Attributes for Natural Ecosystems Affected by O3
Category
Landscape Condition
• Habitat Types
Biotic Condition
• Ecosystems and
Communities
Community Extent
Community
Composition











Trophic Interactions
Insects






Diseases











Community
Dynamics






Species



Mixed conifer forest


Pinus ponderosa

Grassland communities

Coastal sage scrub

Early successional plant
community
Populus tremuloides
& Betula papyrifera
Pinus ponderosa
Pinus taeda


Pinus ponderosa
Pinus ponderosa

Populus tremuloides

Populus tremuloides

Populus hybrids
Populus hybrids
Populus tremuloides

Populus tremuloides

Picea abies &
Picea sitchensis
Pinus ponderosa
Pinus taeda
Pinus sylvestris/
mycorrhizae
Pinus ponderosa/Abies
concolor/calocedrus
decurrens
Populus tremuloides
Pinus ponderosa/
Elymus glaucus
Pinus taeda/diverse
community
Condition Measures



Community composition,
Stand structure

Relative abundance

Species composition

Species cover, richness,
equitability
Species richness, diversity,
evenness
Soil microbial community

Soil microbial community
Fungal morphotypes


Bark beetle severity
Bark beetle productivity and
predator/parasitoid density
Blotch leaf miner
performance
Aphid/natural enemy
abundance
Septoria occurrence
Rust occurrence
Rust occurrence

Forest tent caterpillar/
paratisoid performance
Needle fungi

Root disease x O3 interactions
Canker dimensions
Disease susceptibility

Abundance


Competitive status
O3 sensitivity

Tree growth

References



Miller etal. (1989);
Miller and McBride (1999)

Miller (1973);
Arbaugh et al. (2003)
Ashmore and Ainsworth (1995);
Ashmoreetal. (1995)
Westman( 1979, 1981)

Barbo etal. (1998)

Phillips et al. (2002)

Scagel and Andersen (1997)
Edwards and Kelly (1992);
Qui etal. (1993)

Cobb etal. (1968)
Dahlsten et al. (1997)

Kopper and Lindroth (2003a)

Percy et al. (2002)

Woodbury et al. (1994)
Beare etal. (1999)

Kamosky et al. (2002);
Percy et al. (2002)
Holton et al. (2003)
Maganetal. (1995)

Fenn etal. (1990)
Carey and Kelly (1994)
Bonello etal. (1993)

Minnichetal. (1995)


McDonald et al. (2002)
Andersen et al. (2001)

Barbo et al. (2002)

January 2005
9-257
DRAFT-DO NOT QUOTE OR CITE

-------
         Table 9-22 (cont'd).  Essential Ecological Attributes for Natural Ecosystems
                                              Affected by O3
 Category
Species
Condition Measures
References
   Species and
   Populations

   Population Size
   Genetic Diversity/
   Population Structure
   Population
   Dynamics
Pinus strobus
Firms ponderosa

Lupinus bicolor
Populus tremuloides

Trifolium repens
Plantago major
Trifolium repens
Plantago major
Mortality
Mortality

% population sensitive
% population sensitive

% population sensitive
% population sensitive
Adaptation
Population changes over time
Kamosky(1981)
Carroll et al. (2003)

Dunn (1959)
Berrang etal. (1986, 1989, 1991)

Heagleetal. (1991)
Reiling and Davison (1992a);
Davison and Reiling (1995);
Lyons etal. (1997)

Heagleetal. (1991)
Davison and Reiling (1995)
 Organism Condition

 • Visible Symptoms
Pinus ponderosa
                         Pinus Jeffreyi
                         Prunus serotina
                         Sassafras albidum
                         Populus nigra,
                         Fraxinus excelsior &
                         Prunus avium
                         Fagus sylvatica
                         Fraxinus americana

                         Grassland species
                         Herbaceous species
                         Asclepias exaltata
                         Rudbeckia laciniata &
                         Verbesina occidentalis
                         Asclepias incarnata
Foliar symptoms
                        Foliar symptoms
                        Foliar symptoms
                         Liriodendron tulipfera      Foliar symptoms
                        Foliar symptoms
                        Foliar symptoms
                        Foliar symptoms
                        Foliar symptoms

                        Foliar symptoms
                        Foliar symptoms
                        Foliar symptoms
                        Foliar symptoms

                        Foliar symptoms
Grulke and Lee (1997);
Arbaughetal. (1998;
Salardino and Carroll (1998);
Temple etal. (1992)
Patterson and Rundel (1995);
Salardino and Carroll (1998);
Grulke et al. (2003b)
Fredericksen et al. (1995, 1996);
Chappelka et al. (1997, 1999a,b);
Hillebrandetal. (1996);
Ghosh etal. (1998);
Leeelal. (1999);
Ferdinand et al. (2000, 2003);

Yuska et al.  (2003)
Somersetal. (1998);
Hillebrand etal. (1996);
Chappelka et al. (1999a)
Chappelka etal. (1999a)
Novak et al. (2003)

Gerosa et al. (2003);
Vollenweider et al. (2003a)

Schaub et al. (2003);
Ferdinand et al. (2003)
Bungener et al. (1999a)
Bergmann et al. (1999)
Chappelka etal. (1997)
Chappelka et al. (2003)

Orendovici et al. (2003)
January 2005
                          9-258
                 DRAFT-DO NOT QUOTE OR CITE

-------
         Table 9-22 (cont'd).  Essential Ecological Attributes for Natural Ecosystems
                                            Affected by O3
 Category
Species
Condition Measures
References
   Physiological Status
   Reproductive Status
Pinus halepensis
Populus tremuloides
Betula pendula
Fagus sylvatica
Populus tremuloides
Pinus ponderosa

Festiva ovina
Betula pubescens
Fagus sylvatica

Populus tremuloides x
P. tremula
Prunus serotina
Fraxinus excelsior
Populus tremuloides
Pinus ponderosa

Betula pendula

Betula pendula
Acer saccharum
Apocynun
  androsaemifolium
Buddleia davidii
Rubus cuneifolius
Plantago major,
Allometry
Crown architecture
Crown architecture
Crown architecture
Root dry weight
Root/shoot ratio

Root/shoot ratio
Root/shoot ratio
Root/shoot ratio

Root/shoot ratio

Leaf area
Leaf area
Leaf area index
Carbon allocation to
  mycorrhizae
Decreased winter bud
  formation
Delayed bud break
Early bud break
Flowering time
Flowering time
Flowering time
Pollen germination &
Pollen tube elongation
                         Fragaria x ananassa     Fruit yield

                         Plantago major          Seed yield



                         Understorey herbs        Seed yield
Welburn and Wellburn (1994)
Dicksonetal. (2001)
Kull et al. (2003)
Stribley and Ashmore (2002)
Colemanetal. (1996)
Grulkeetal. (1998);
Grulke and Balduman, (1999)
Warwick and Taylor, (1995)
Mortensen(1998)
                                                                             Paludan-Muller et al. (1999);
                                                                             Landollelal. (2000)
                                                                             Oksanenetal. (200la)
Neufeld et al. (1995)
Wiltshire et al. (1996)
Karnosky et al. (2003a)
Andersen and Rygiewicz (1995)
Oksanen (2003a,b)

Prozherina et al. (2003)
Bertrand et al. (1999)

Bergweiler and Manning,
(1999)
Findleyetal. (1997)
Chappelka (2002)
Stewart (1998)
                                                     Drogoudi and Ashmore
                                                     (2000,2001)
                                                     Reiling and Davison (1992b);
                                                     Pearson et al. (1996); Whitfield
                                                     et al. (1997); Lyons and Barnes
                                                     (1998)
                                                     Harward and Treshow (1975)
January 2005
                         9-259
                 DRAFT-DO NOT QUOTE OR CITE

-------
         Table 9-22 (cont'd).  Essential Ecological Attributes for Natural Ecosystems
                                            Affected by O3
 Category
Species
Condition Measures
References
 Ecological Processes
 • Energy Flow

   Primary Production     Pinus ponderosa
                         Pinus ponderosa
                         Populus tremuloides
                         Betula pendula
                         Betula pendula

                         Quercus rubra
                         Populus tremuloides
                         Populus tremuloides
                         Pinus ponderosa
                         Quercus rubra

                         Populus tremuloides
                         Pinus ponderosa

                         Betula pendula
                         Betula pendula
                         Fragaria vesca
                         Pinus taeda
                         Lespedeza cuneata &
                         Schizacbyrium
                         scoparium
                         Liriodendron tulip/era
                         Prunus serotina
                         Pinus jeffreyi
                         Pinus ponderosa
                         Pinus strobus
                         Pinus taeda
                       Photosynthesis
                       Needle retention
                       Photosynthesis
                       Photo sy nthe sis/conductance
                       Stem respiration & radial
                         growth
                       Root turnover
                       Soil respiration
                       Soil respiration
                       Soil respiration
                       Carbon partitioning &
                         allocation
                       Carbon allocation
                       Carbon allocation

                       Carbon allocation
                       Carbon allocation
                       Carbon allocation
                       Root respiration
                       Yield
                       Radial growth
                       Radial growth
                       Radial growth
                       Radial growth (no effect)
                       Radial growth
                       Radial growth
                             Miller etal. (1969);
                             Clark etal. (1995);
                             Takemoto et al. (1997);
                             Grulke et al. (2002b)
                             Temple etal. (1993)
                             Colemanetal. (1995b);
                             Noormets et al. (2001a,b);
                             Sharmaetal. (2003;
                             Karnosky et al. (2003a)
                             Oksanen (2003a,b)
                             Matyssek et al. (2002)

                             Keltingetal. (1995)
                             Colemanetal. (1996)
                             King etal. (2001)
                             Andersen and Scagel (1997);
                             Scagel and Andersen (1997),
                             Andersen (2000)
                             Samuelson and Kelly (1996)
                             Colemanetal. (1995b)
                             Andersen et al. (1997);
                             Grulke etal. (1998);
                             Grulke and Balduman (1999);
                             Grulke etal. (2001)

                             Karlsson  el al. (2003)
                             Oksanen and Saleem, (1999);
                             Salem etal. (2001)

                             Manninen et al. (2003)
                             Edwards (1991)
                             Powell et al. (2003)

                             Somers etal. (1998)
                             Vollenweider et al. (2003b)
                             Peterson etal. (1987)
                             Peterson etal. (1993)
                             Bartholomay et al. (1997)
                             and Downing (1995,  1996)
                             Braunetal. (1999)
Fagus sylvatica
Picea abies
Populus tremuloides
Pinus ponderosa
Stem volume
Stem volume
Volume growth
Root growth
Wallm et al. (2002)
Isebrands et al. (2001)
Andersen et al. (1991)
January 2005
                         9-260
                DRAFT-DO NOT QUOTE OR CITE

-------
         Table 9-22 (cont'd).  Essential Ecological Attributes for Natural Ecosystems
                                             Affected by O3
 Category
Species
Condition Measures
References
   Net Ecosystem
   Production
   Material Flow
   Organic Carbon
   Cycling
   Nitrogen Cycling
   Other Nutrient
   Cycling
Northern hardwoods
Northern hardwoods
   Growth Efficiency      Plantago major
                         Grassland species
                         Native herbs

                         Grasses and herbs

                         Populus tremuloides
                         Fagus sylvatica
                         Picea abies
                         Primus serotina
Populus tremuloides &
Be tula papyri/era
Andropogon virginicus
& Rubus cuneifolius
Liriodendron tulipera

Populus deltoides
Pinus ponderosa
Pinus sylvestris
Pinus ponderosa
Pinus taeda

Prunus serotina &
Liriodendron tulip/era

Picea sitchensis &
Pinus sylvestris
Pinus ponderosa
 Hydrology and Geomorphology

 • Water Budget          Picea rubens

                         Pinus armandi
                         Pinus jeffreyi
                         Picea abies

                         Fraxinus excelsior
                         Be tula pendula

                         Populus hybrids
NPP estimates
Biomass estimates

Relative growth rate
                        Relative growth rate
                        Relative growth rate

                        Relative growth rate

                        Relative growth rate
                        Relative growth rate
                        Relative growth rate
                        Relative growth rate
Altered foliar C:N ratio and
  N resorption efficiency
Litter decomposition rate

Litter decomposition rate

Litter decomposition rate
Litter decomposition rate
Litter decomposition
  (no effect)

Altered foliar N
Foliar N & O3 exposure
 (no effects)
Altered foliar N metabolism
Altered foliar N

Foliar leaching (no effect)

Nutrient availability & O3
                        Water-use efficiency
                          (no effect)
                        Water-use efficiency
                        Canopy transpiration
                        Transpiration, xylem sap flow

                        Stem flow of water
                        Water-use efficiency

                        Water-use efficiency
Laurence et al. (2000)
Hogsettetal. (1997)

Reiling and Davison (1992a);
Davison and Reiling (1995);
Lyons et al. (1997);
Davison and Barnes (1998)

Bungener et al. (1999b)
Warwick and Taylor (1995)

Pleijel and Danielsson (1997)

Yun and Laurence (1999)
Bortier et al. (2000)
Karlsson et al. (2002)
Lee et al. (2002)
Lindroth et al. (2001)

Kim etal. (1998)

Scherzer and Rebbeck (1998)
Findlay and Jones (1990)
Fenn and Dunn (1989)
Kainulainen et al. (2003)
Momen and Helms (1996)
Bytnerowicz et al. (1990)

Mandersheid et al. (1992)

Boerner and Rebbeck (1995)
Skeffington and Sutherland
(1995)
Bytnerowicz et al. (1990)
                              Laurence et al. (1997)

                              Shan etal. (1996)
                              Grulke et al. (2003a)
                              Maier-Maercker (1997)

                              Wiltshire et al. (1994)
                              Maurer and Matysek (1997)

                              Reich and Lassoie (1984)
January 2005
                          9-261
                 DRAFT-DO NOT QUOTE OR CITE

-------
                Table 9-22 (cont'd).  Essential Ecological Attributes for Natural Ecosystems
                                                Affected by O3
         Category
Species
Condition Measures
References
         Natural Disturbance Regimes

         • Frequency           Pinus ponderosa
           Intensity
           Extent

           Duration
Pinus ponderosa


Picea sitchensis
Pinus halepensis

Picea rubens
Fagus sylvatica


Picea abies
Pinus ponderosa

Pinus ponderosa

Pinus ponderosa
                     Frequency of fire
Occurrence of bark beetle
 outbreaks

Winter damage
Reduced winter damage

Freezing tolerance
Drought stress
Drought stress



Fire intensity

Extent of bark beetle attack

Duration of bark beetle attack
McBride and Laven (1976);
Minnichetal. (1995);
Miller and McBride (1999)
Pronosetal. (1999);
Dahlsten et al. (1997)

Lucas etal. (1988)
Wellburn and Wellburn, (1994)

Waiteetal. (1994)
Pearson and Mansfield
(1993, 1994)

Maier-Maercker (1998);
Maier-Maercker and
Koch (1992)

Miller and McBride (1999)

Minnichetal. (1995)

Minnichetal. (1995)
         Source: Young and Sanzone (2002.
 1      conductance are greatest and these factors enhance O3 uptake and injury. Decline of A. religiosa
 2      (oyamel) in the Desierto de los Leones National Park is a well-known example of dramatic
 3      dieback and mortality of entire forest stands due primarily to air pollution stress

 4      (Alvarado-Rosales and Hernandez-Tejeda, 2001).  Other factors, such as a lack of stand
 5      thinning, also contribute to forest decline.  Lead in automobile gasoline was phased out in 1990,
 6      and foliar concentrations of heavy metals in forest species are not now at phytotoxic levels (Fenn
 7      et al. 2001a).  Sulfur dioxide concentrations decreased in the early 1990s as a result of regulatory

 8      mandates limiting their emissions.  Sensitive plants in the NE and NW sectors of the Mexico
 9      City urban zone where concentrations are highest may still be impacted by  exposure to ambient
10      SO2 levels.  Deposition of ionic forms of N and S are high in forested areas southwest of Mexico
11      City.  The effects of these chronic nutrient inputs to the forest are only beginning to be

12      investigated and understood.
        January 2005
                       9-262
               DRAFT-DO NOT QUOTE OR CITE

-------
 1           The ecological perturbations caused by severe air pollution exposures in forests located
 2      downwind of Mexico City are expected to continue for the near future (the next 5 to 10 years),
 3      largely as a result of high O3 concentrations, as well as of N oxides emissions. The longer-term
 4      response is more uncertain, and depends largely on the effectiveness of regulatory emissions
 5      control strategies.  Currently, pollutant levels are declining. Forest responses to this trend will
 6      depend on how long it takes to reduce levels sufficiently to allow sensitive species to recover.
 7      Some of the change to the ecosystem is probably irreversible, such as the loss of lichen diversity
 8      and of other O3-sensitive species (Zambrano et al., 2002).
 9
10      9.7.2.2 San Bernardino Mountains
11           The San Bernardino Mountains lie east of the Los Angeles Air Basin (California, USA),
12      and significant levels of pollution have been transported into the mountain range, including a
13      Class I Wilderness area. The effects of O3 exposure on the mixed conifer forest of the
14      San Bernardino Mountains is perhaps the  longest and most thoroughly documented O3
15      ecological effects evaluation (Miller and McBride, 1999).  In this classic case study linking
16      tropospheric O3 exposure to damage to an entire forest ecosystem (U.S. Environmental
17      Protection Agency, 1996) (Table 9-23), Miller et al. (1963) first identified the unique foliar
18      chlorotic mottle that was occurring on two of the dominant tree species, Pinus ponderosa and
19      P.jeffreyi. Levels of O3 averaging 100-120 ppb over 24 hours with 1-hour peaks well into the
20      200 ppb range were common in the region in the 1960s and 1970s (Miller and McBride, 1999).
21      Single hour peak values have declined in recent years  due to heavily regulated pollution control,
22      but O3 concentrations in the moderate range continue to rise and accumulate, increasing the
23      overall cumulative exposure (Arbaugh et al.,  1998; Takemoto et al., 2001; Lee et al., 2003).
24           Since the 1996 O3 AQCD (U.S. Environmental Protection Agency, 1996) was written, the
25      concurrent role of N deposition in modifying ecosystem response to O3 exposure in the San
26      Bernardino Mountains has been further elucidated (Fenn et al., 1996, 2003; Bytnerowicz et al.,
27      1999; Takemoto et al., 2001; Bytnerowicz, 2002).  Both O3 exposure and N deposition reduce
28      foliar retention (Grulke and Balduman,  1999) and alter tissue chemistry of both needles and litter
29      (Poth and Fenn,  1998). In addition, confounding factors such as drought and fire suppression
30      add to the complexity of ecosystem response (Minnich et al., 1995; Takemoto et al., 2001;
31      Arbaugh et al., 2003). Extensive crown injury measurements have also been made, linking

        January 2005                             9-263       DRAFT-DO NOT QUOTE OR CITE

-------
             Table 9-23. Case Studies Demonstrating the Ecological Effects of O3
 Study
Keystone
Species
                Period
 Study Type    Studied   Key Ecological Findings
 Valley of
 Mexico
Pinus hartwegii,
Abies religiosa
Field transects
 San Bernardino   Pinus ponderosa,    Field transects
 Mountains        P. jeffreyi
35 yrs    • Significant foliar injury
           (de Bauer, 1972; Krupa and Bauer, 1976;
           Hernandez-Tejeda, 1986)
         • Community composition changes
           (Alvarado-Rosales and Hernandez-
           Tejeda, 1986)
         • Species richness changes
           (Zambrano et al., 2002)

40 yrs    • Community composition changes
           (Miller, 1973; Minnichet al., 1995;
           Arbaughetal.,2003)
         • Population changes
           (McBride and Laven, 1999)
         • O3/pine^ark beetle interaction
           (Pronosetal. 1999)
         • Altered C flows
           (Grulke et al., 2002b; Grulke et al., 1998;
           Grulke andBalduman, 1999; Grulke
           et al., 2001; Arbaugh et al., 1999)
         • Interaction of O3, drought, N deposition:
           (Fennetal., 1996; Grulke, 1999;
           Takemoto et al., 2001)
         • Altered carbon cycling
           (Arbaugh et al., 1999)
 Sierra Nevada
 Mountains
Pinus ponderosa,
P. jeffreyi
 Appalachian
 Mountains
Fraxinus
americana,
Liriodendron
tulip/era,
Pinus strobus,
Prunus serotina
    Field        35 yrs    • Wide-scale nature of effects (Miller and
                           Millecan, 1971; Miller etal,. 1972)
                          • Link to decreased growth
                           (Peterson et al., 1987, 1991, 1995)
                          • Quantification of O3 flux
                           (Bauer et al., 2000; Panek et al., 2002;
                           Goldstein etal., 2003)
                          • Cumulative O3 effects
                           (Takemoto et al., 1997)
                          • Canopy level responses
                           (Grulke etal., 2003a,b)
                          • Population changes (Carroll et al., 2003)

    Field        25 yrs    • Link of visible symptoms to growth
                           decreases (McLaughlin et al.,  1982;
                           Somersetal., 1998)
                          • Wide-scale nature of effects (Chappelka
                           et al., 1999a; Hildebrand et al., 1996)
January 2005
                              9-264
                          DRAFT-DO NOT QUOTE OR CITE

-------
               Table 9-23 (cont'd). Case Studies Demonstrating the Ecological Effects of O3
Study
Aspen FACE











Keystone
Species
Acer sacchamm,
Betula
papyri/era,
Populus
tremuloides







Period
Study Type Studied Key Ecological Findings
Open-air O3 6 yrs • Competitive interactions
exposure (McDonald et al., 2002)
• O3/aspen/rust interaction
(Karnoskyetal, 2002)
• Plant- insect interactions (Percy et al.,
2002; Holtonetal, 2003)
• C & N cycling (Lindroth et al., 2001;
King etal., 2001)
• Moderation of CO2 responses by O3
(Isebrands et al., 2001; Wustman et al.,
2001; McDonald et al., 2002;
Karnoskyetal.,2003a)
         Plantago        Plantago major        Field       20 yrs   • Population structure (Davison and
                                                                Reiling, 1995; Lyons et al., 1997)
                                                              • O3 resistance
                                                                (Reiling and Davison, 1992a)
                                                              • Adaptation (Davison and Reiling, 1995)
         Carpathian      Pinus sylvestris,        Field       15 yrs   • Significant foliar injury
         Mountains      Picea abies                              • Community composition changes
                                                              • Species diversity changes
 1      ambient O3 exposure data to chlorotic mottle and fascicle retention (Arbaugh et al., 1998).
 2      Ozone exposure and N deposition reduce carbon allocation to stems and roots (Grulke et al.,
 3      1998; 2001), further predisposing trees to drought stress, windthrow, root diseases, and insect
 4      infestation (Takemoto et al., 2001). Increased mortality of susceptible tree species (ponderosa
 5      and Jeffrey pine, Pinus ponderosa, P. Jeffreyii) has shifted community composition towards
 6      white fir and incense cedar (Abies concolor, Libocedrus decurrens) and has altered forest stand
 7      structure (Miller et al., 1989). Ozone exposure is implicated in projected changes in stand
 8      composition (McBride and Laven, 1999) toward a predominance of oaks, rather than mixed
 9      conifer forests. Forest understory species have also been affected (Temple, 1999). These
10      individual species responses collectively have affected trophic structure and food web  dynamics
11      (Dahlsten et al., 1997; Pronos et al., 1999), as well as C and N cycling (Arbaugh et al., 2003)
12      (Table 9-24).  Because of the high N deposition in the San Bernardino Mountains, it is difficult
13      to separate out effects of only O3 versus those due to the combined effects of O3 and N
14      deposition (Table 9-25).
        January 2005
9-265
DRAFT-DO NOT QUOTE OR CITE

-------
   Table 9-24.  The Most Comprehensively Studied Effects of O3 on Natural Ecosystem is
      the San Bernardino Mountain Forest Ecosystem. Citations Focus on Research
                             Published Since U.S. EPA (1996).
 Pollutant Occurrence
    Reference
 O3 exposures N deposition
 Cellular, Biochemical

 Foliar pigments
 Antioxidants
 Foliar Responses

 Foliar Symptoms


 Gas Exchange

 Photosynthesis & Conductance
 O3flux
 Foliar nutrients

 Whole Organism

 Growth/Biomass
 • Aboveground
 • Belowground
 • Root/shoot ratio
 • Carbon allocation
 • Crown vigor


 Ecosystem

  Community dynamics/succession
  Simulations

  Understory vegetation

 Pest interactions
 • Bark beetle/predators
 • Disease occurrence
 • Litter decomposition

 Disturbance

 • Bark beetle occurrence
 • Fire frequency
    Grulke et al. (1998); Fenn et al. (1996, 2000,
    2003); Bytnerowicz et al. (1999)
    Grulke and Lee (1997); Grulke (1999); Tausz
    et al. (1999a,b 2001); Tausz et al. (1999a,b,c,
    2001)
    Grulke and Lee (1997); Arbaugh et al. (1998);
    Miller and Rechel (1999)
    Grulke (1999); Grulke and Retzlaff (2001);
    Grulke et al. (2002a,b)
    Grulke and Balduman (1999)
    Grulke et al. (1998); Grulke and Balduman
    (1999)
    Grulke and Balduman (1999)
    Grulke etal. (2001)
    Arbaugh et al. (1998); Miller and Rechel (1999)
    McBride and Miller (1999); Arbaugh et al. (2002)
    Arbaugh etal.  (1999)

    Temple (1999)
    Dahlsten et al. (1997); Pronos et al. (1999)
    Miller and Rechel (1999); Pronos et al. (1999)
    Minnich et al. (1995); Minnich (1999); Miller and
    McBride (1999)
January 2005
9-266
DRAFT-DO NOT QUOTE OR CITE

-------
  Table 9-25. Effects of Ozone, Ozone and N Deposition, and Ozone and Drought Stress
    on Pinus ponderosa and Pinus jeffreyi in the Sierra Nevada and the San Bernardino
          Mountains, California. Citations are Focused on Research Published
                              Since U.S. EPA (1996).
Foliar Biochemistry and
Tissue Chemistry
Total ascorbate
Dehydroascorbate
Total glutathione
Oxidized glutathione
a Carotenoids
Foliar nitrogen
C:N ratio of foliage2
Starch
Chlorophyll content
Gas Exchange
A,,,^ lower canopy
Amax whole canopy
A^ seedlings
Stomatal limitation
Stomatal conductance
Foliar respiration
O3flux
O3 O3 + N deposition
d1 d
i n.d.
d i
i i
i n.d.
d i
i n.d.
ad. d
d id

n.d. i
d ad.
di n.d.
a to ps n.d.
d di
as. i
d as.
O3 + Drought
i
d
d
d
d
d
d
i
d

d
d
ad.
i
d
d
d
References
Tausz et al. (2001); Grulke et al.
(2003b)
Grulke et al. (2003b)
Tausz etal. (2001)
Tausz etal. (2001)
Grulke et al. (2003b)
Grulke and Lee (1997);
Grulke etal. (1998);
Poth and Fenn (1998)
Poth and Fenn (1998);
Grulke et al. (2003b)
Grulke etal. (2001)
Takemoto et al. (1997); Grulke
and Lee (1997; Grulke et al.
(1998, 2003b); Grulke (1999);
Tausz etal. (2001)

Grulke (1999); Grulke and
Retzlaff (2001); Grulke et al.
(2002b); Panek (2004)
Panek and Goldstein (2001);
Grulke et al. (2003b); Misson
et al. (2004)
Grulke and Retzlaff (2001)
Panek and Goldstein (2001)
Grulke (1999; Grulke et al.
(2003a); Panek (2004)
Grulke et al. (2002a; Grulke
(1999)
Panek and Goldstein (2001);
Panek et al. (2002; 2003); Grulke
et al. (2002b; 2004)
January 2005
9-267
DRAFT-DO NOT QUOTE OR CITE

-------
             Table 9-25 (cont'd). Effects of Ozone, Ozone and N Deposition, and Ozone and
            Drought Stress on Pinus ponderosa and Pinus jeffreyi in the Sierra Nevada and the
               San Bernardino Mountains, California. Citations are Focused on Research
                                    Published Since U.S. EPA (1996).
Growth and Productivity
Foliar biomass
Height growth
Bole diameter growth
Fine root biomass
03
ad.
ad.
d
d
O3 + N deposition
i
i
i
d
O3 + Drought
d
d
d
i
References
Grulke and Balduman (1999)
Grulke and Balduman (1999)
Grulke and Balduman (1999)
Grulke etal. (1998)
        Leaf Surfaces

        Stomatalocclusion           i            n.d.              n.d.       Bytnerowiczetal. (1999);
                                                                       Bytnerowicz and Turunen (1994)

        Trophic Interactions

        Bark beetle                n.s.            i                i        Pronos etal. (1999)

        Ecosystem Level

        Competitive indices         n.d.            d                i        Miller and Rechel (1999)


        'Responses are shown as significant increases (i), significant decreases (d), both significant decreases and
         increases reported (di), nonsignificant effects (n.s.), and no data (n.d.) compared to trees or seedlings at field sites
         with lower ozone, drought stress, or lack of significant N deposition (< 10 kg ha"1 yf'). Frequently n.d. was used
         for lack of a control site without compounding high N deposition. Foliar analyses and leaf surface properties
         were largely determined from previous year needles. Gas exchange data were generally from previous year
         needles at peak growing season, prior to late summer drought (mid to late July).
        Abbreviations:  C = carbon; N = nitrogen; A^ = maximum photosynthesis rate.
1      9.7.2.3  Sierra Nevada Mountains

2           The western slope of the Sierra Nevada Mountains in central and southern California has

3      also been exposed to elevated O3 for a long time, although the effects have been much less than

4      those observed in the San Bernardino Mountains. Symptoms of O3 injury have been found on

5      ponderosa and Jeffrey pines in all of the Sierra Nevada National forests and parks (Carroll et al.,

6      2003). First identified as a problem in the 1970s (Miller et al., 1972), elevated O3 with daytime

7      means of 60-80 ppb are common (Bohm et al., 1995; Bauer et al., 2000; Bytnerowicz et al.,

8      2002a; Panek et al., 2002). The west-slope Sierra Nevada forests are also exposed to a wide

9      range of additional gaseous and paniculate pollutants, including various S and N compounds


       January 2005                               9-268        DRAFT-DO NOT QUOTE OR CITE

-------
 1      (Bytnerowicz et al., 1999; Takemoto et al., 2001; Fenn et al., 2003), but at levels much lower
 2      than in the San Bernardino Mountains. Typical O3-induced visible foliar symptoms, including
 3      chlorotic mottle, chlorophyll degradation, and premature senescence, are  commonly found on
 4      sensitive genotypes ofponderosapine (Peterson et al., 1991; Arbaugh et  al., 1998; Staszak et al.,
 5      2003) and Jeffery pine (Peterson et al., 1987; Patterson and Rundel, 1995; Arbaugh et al., 1998;
 6      Grulke et al., 2003b). Other important conifers in the region, such as giant sequoia, appear to be
 7      relatively tolerant to O3 (Grulke et al., 1996).  The symptoms of foliar injury and growth
 8      reductions have been verified on seedlings in O3 exposure chambers (Temple, 1988; Momen and
 9      Helms, 1996; Momen et al., 2002).
10           Ozone foliar injury of dominant pine species in the Sierra Nevada Mountains is correlated
11      to decreased radial growth in both ponderosa pine (Peterson et al., 1991)  and Jeffreyipine
12      (Peterson et al., 1987; Patterson and Rundel, 1995).  Because of the large amount of intraspecific
13      variation in O3 sensitivity in these two species, O3 exposure may be a selective agent (Patterson
14      and Rundel, 1995), with differential mortality rates for sensitive individuals (Carroll et al.,
15      2003). The region's forests may also be experiencing subtle changes in species composition and
16      community structure (Patterson and Rundel, 1995; Takemoto et al., 2001).
17           Based on fire scar dating, reconstructions of stand age classes, historical records, and
18      present stand structure, fire has been largely excluded in western forests for the last 75 to 100
19      years (Minnich, 1999; Minnich and Padgett, 2003).  Fire exclusion has  resulted in fewer, large,
20      stand-replacing fires rather than a mosaic of smaller, lower intensity fires. The change in fire
21      intensity may have selectively altered stand structure, fitness and competitiveness of component
22      species,  and their susceptibility to atmospheric pollutants and other stressors (Minnich, 1999).
23      Short-lived (50-80 years)  species such as knobcone and Coulter pine, which occur at the
24      interface between the chaparral and the mixed conifer forest, may already have been selected for
25      O3 tolerance due to seedling establishment (the most sensitive tree age class in conifers) after
26      large fires in the 1950's (Minnich, 1999).  Strong measures to suppress fires have largely kept
27      chaparral fires from invading the mixed conifer forests, and stand densification in the mixed
28      conifer zone has increased.  High stand density, in turn, may weaken the younger cohorts and
29      increase sensitivity to atmospheric pollution (Minnich, 1999).
30           Other disturbances that play a potential role in sensitivity to atmospheric pollution include
31      cycles of drought stress.  Nearly every decade is marked by one or more years of very low

        January  2005                             9-269        DRAFT-DO NOT QUOTE OR CITE

-------
 1      precipitation (Graumlich, 1993).  During extended periods of drought, foliar injury is lower than
 2      in subsequent years with higher average precipitation (Carroll et al., 2003). In the first several
 3      years (1975-1977) of a Sierran-wide assessment of O3 injury to pines, O3 injury increased,
 4      because of greater water availability due to greater stomatal conductance and, presumably,
 5      greater O3 uptake. Trees instrumented with monitors to directly measure canopy transpiration
 6      had 20% greater stomatal conductance in mesic microsites (riparian areas, mid-slope seeps) than
 7      trees in xeric microsites (rock outcrops) (Grulke et al., 2003a). Although the Sierra Nevada
 8      experienced a prolong drought between 1987 and 1993, it was less severe and O3 injury did not
 9      significantly decrease (Carroll et al., 2003).  The same plots showed only a slight increase in O3
10      injury between 1993 and 2000. Drought stress, in general, can make trees more susceptible to
11      insect and pathogen infestation. However, serious outbreaks of insect infestation are believed to
12      be indicators, not a cause, of existing stress in the forest (Wickman, 1992).
13
14      9.7.2.4 Appalachian Mountains
15           The southern Appalachian Mountain region experiences some of the highest O3 exposures
16      of any natural areas in the eastern United States (Mueller, 1994; Hildebrand et al., 1996;
17      Samuelson and Kelly, 1997; Chappelka et al., 1997). Since the region is the home of the
18      Shenandoah and Great  Smokey Mountains National Parks, which have Class I air quality
19      designations by the 1977 Clean Air Act, there has been considerable study of the region's
20      dominant forest species to determine O3 effects. Visible foliar symptoms of O3 have been found
21      in natural ecosystems consisting of Sassafras albidum (Chappelka et al., 1999a), Prunus serotina
22      (Hildebrand et al., 1996; Chappelka  et al., 1997; 1999b; Samuelson and Kelly, 1997),
23      Liriodendron tulipifera and Fraxinus Americana (Hildebrand et al., 1996; Chappelka et al.,
24      1999a). Visible foliar symptoms induced by O3 have been recreated on the same species in
25      chamber studies (Duchelle et al.,  1982; Chappelka et al., 1985; Chappelka and Chevone, 1986;
26      Samuelson, 1994; Fredericksen et al., 1995).  No response to O3 exposure has been found for
27      other hardwood trees, nor for the three conifer species tested (Neufeld et al.,  2000).
28           Long term foliar injury symptoms have been correlated with decreased radial growth in
29      tulip poplar and cherry {Liriodendron tulipifera and Prunus serotina, Somers et al. (1998) and
30      with decreased biomass in cherry (Neufeld et al., 1995).  Although climatic conditions (drought)
31      largely explained radial growth reductions, O3 exposure may have also contributed (McLaughlin

        January 2005                             9-270       DRAFT-DO NOT QUOTE OR CITE

-------
 1      and Downing, 1996).  Ozone exposure may also be affecting the understory vegetation in the
 2      region (Duchelle and Skelly, 1981; Duchelle et al., 1983; Chappelka et al., 1997; Chappelka
 3      et al., 2003; Davison et al., 2003) and community composition (Barbo et al., 1998), through
 4      impacts both on growth and reproduction (Chappelka, 2002).  Foliar litter from trees exposed to
 5      elevated O3 have lower decomposition rates (Kim et al. 1998). Other air pollutants are likely to
 6      be found in this ecosystem but not at such high deposition values found in the California studies.
 7      A decline in forest health in the northern Appalachians has been primarily attributed to the
 8      effects of acidic fog and rain on soil acidification, lower Ca2+ availability, reduction in fine root
 9      biomass, and modification of cuticular wax. However, fog- and O3-exposed red spruce forests
10      also show winter injury (Percy, 2002).
11
12      9.7.2.5   Plantago Studies in the United Kingdom
13           One of the most well-documented studies of population and community response to O3
14      effects are the long-term studies of Plantago major in native plant communities in the United
15      Kingdom (Reiling and Davison, 1992a; Davison and Reiling,  1995; Lyons et al., 1997).
16      Sensitive populations of P. major had significant growth decreases in elevated O3 (Reiling and
17      Davison, 1992b and 1992c; Pearson et al., 1996; Whitfield et al., 1997)  and reduced fitness as
18      determined by decreased reproductive success (Reiling and Davison, 1992b; Pearson et al.,
19      1996). While spatial comparisons of population responses to O3 are complicated by other
20      environmental factors, rapid changes in O3 resistance were imposed by ambient levels and
21      variations in O3 exposure (Davison and Reiling, 1995). Molecular patterns of genetic variation
22      suggest that change in O3 resistance over time probably resulted from selection on genotypes
23      already present in local populations, rather than through  an influx of new P. major germplasm
24      (Wolff et al., 2000).  The highest correlations between O3 resistance and ambient O3
25      concentrations occurred at the  site of seed collection (Lyons et al., 1997), rather than between O3
26      resistance and other climatic variables, as was found for  aspen (Berrang et al.,  1991).
27
28      9.7.2.6   Forest Health in the  Carpathian Mountains
29           The Carpathian Mountains cross five countries (the Czech Republic, the  Slovak Republic,
30      Poland, Romania, and the Ukraine) and contain many national parks and several biosphere
31      reserves. The forests were largely cleared in the 15th century, and were then planted in Norway

        January 2005                              9-271       DRAFT-DO NOT QUOTE OR CITE

-------
 1      spruce plantations.  As elevation increases, beech (Fagus sylvaticd) or beech-fir (Abies alba)
 2      forests grade into Norway spruce (Picea abies) or spruce-fir forests. Near the treeline, Norway
 3      spruce combines with dwarf mountain pine (Pinus mugo). Dwarf mountain pine forms an
 4      almost pure stand just below alpine vegetation.
 5           The forests of the Carpathian Mountains have been subjected to anthropogenic stressors
 6      (shepherding, metal mining, wood harvest for structures and paper) for hundreds of years, as
 7      described for the Tatra Mountains in the southern Carpathians (Wezyk and Guzik, 2004). The
 8      Carpathians have been subjected to regional air pollution stresses since industrialization.  Most
 9      of the effects of air pollution on forest health degradation were due to (1) heavy metal
10      deposition, (2) soil  acidification by acid deposition, and (3) subsequent pest outbreaks, the
11      combination of which led to the forest decline and dieback between 1970 and 1989 (Dunajski,
12      2004). By the 1980's, industrial pollutants such as SO2 and heavy metals significantly declined,
13      but O3 exposure has continued to increase (Bytnerowicz et al., 2004).  Increased ownership and
14      use of private cars in Central Europe, as well as long-range transport of O3 from western Europe,
15      are believed to be responsible for the continued increase in photooxidants.  In 1995, drought
16      resulted in significant mortality, as well as an epidemic of bark beetle infestation in subsequent
17      years. A network of air quality monitoring sites was installed across Europe in the late 1980's, as
18      part of the International Cooperative Programme on Assessment and Monitoring of Air Pollutant
19      Effects on Forests (ICP Forests). Mean defoliation rates for six important forest species across
20      Europe have increased or remained unchanged from 1989 to 1999 (Percy, 2002).  Ozone
21      concentrations experienced in the Tatra Mountains, especially along the  southern  slopes,
22      occasionally reach  190-200 ppb as two-week-long averages, with the highest values experienced
23      in early summer at  elevations of 1700 to 2300 m (Bytnerowicz et al., 2004). In other parts of the
24      Carpathian Mountains, peak two-week averages of O3 concentrations were lower, at  160 ppb
25      (Bytnerowicz et al., 2002). For all trees inventoried, about 13% exhibited greater than 25%
26      defoliation during 1997 to 2000. There was no difference between extent of damage for
27      broadleaves or conifers.  Trees in Poland and the Czech Republic are the most affected by air
28      pollution, and the least damaged forests are in Romania (Badea et al., 2002).
29           The extent to  which O3 exposure affects forest health degradation,  and slows forest
30      degradation, is still unknown in Europe. In many of the published studies, response to a known
31      O3 gradient is largely confounded by other pollutants and/or climatic gradients (Bytnerowicz

        January 2005                              9-272        DRAFT-DO  NOT QUOTE OR CITE

-------
 1      et al., 1996; Szaro et al., 2002; Widadki et al., 2004). Current levels of ambient O3 are believed
 2      to be high enough to reduce bole radial growth (Percy, 2004). Although average O3
 3      concentration alone was not related to bole growth, the peak hourly O3 concentration was
 4      negatively correlated to growth (Muzika et al., 2002).  Recent evidence indicates that canopy
 5      health of European white oak (Quercus robur\ Norway spruce, maritime pine (Pinus pinaster).,
 6      and beech has significantly declined (Huttunen et al., 2002).  However, the canopy health of
 7      Scots pine (Pinus sylvestris) has improved. The network of air quality monitoring stations and
 8      forest plots is extensive and active. Subsequent correlative analyses including both
 9      meteorological and air quality attributes throughout the EU will help to determine the specific
10      role of O3 exposure in forest decline.  Historical effects of anthropogenic disturbance may still
11      be confounding.
12
13      9.7.2.7  Field Exposure System (FACE), Rhinelander, Wisconsin
14          The Aspen Free-Air CO2 Enrichment (FACE) facility was designed to examine the effects
15      of both elevated CO2 and O3 on aspen (Populus tremuloides\ birch (Betula papyfera), and sugar
16      maple (Acer saccharum) in a  simple reconstructed plantation characteristic of Great Lakes
17      Aspen-dominated forests (Karnosky et al.,  1999, 2003a). Instead of using chambers to expose
18      the plants to desired  gas concentrations, the gas is piped up vertical delivery tubes in the open
19      air.  The vertical delivery pipes surround a 30-m diameter circular plot with five different aspen
20      clones in half of the plot, one  quarter of the plot planted in aspen and birch, and one quarter in
21      aspen and maple.  The O3 treatment for the first five years was 1.5* ambient, with ambient O3
22      exposures averaging 35 to 37  ppb (12 h daytime over the growing season) compared to elevated
23      O3 rings averaging 49 to 55 ppb for the same time period (Karnosky et al., 2003a).
24          Elevated CO2, elevated O3, and elevated CO2  + O3 have had effects on most system
25      components being measured in the study (Table 9-26; Karnosky et al., 2003a).  One interesting
26      finding of the project has been the nearly complete offset by elevated O3 of the enhancements
27      induced by elevated  atmospheric CO2 for the pioneer keystone species P. tremuloides (Isebrands
28      et al., 2001) and B. papyri/era (Percy et al., 2002), even though O3 exposure alone did not
29      always result in a significant response when compared to controls. They also found evidence
30      that the effects on above- and below-ground growth and physiological processes have cascaded
31      through the ecosystem, even affecting microbial communities (Larson et al., 2002; Phillips et al.,

        January 2005                             9-273       DRAFT-DO NOT QUOTE OR CITE

-------
       Table 9-26. Summary of Responses of Populus tremuloides to Elevated CO2
    (+200 fimol mol"*), O3 (1.5 x ambient), or CO2+O3 Compared with Control During
                Three Years of Treatments at the Aspen FACE Project
                       (Modified from Karnosky et al. 2003a)
Foliar Gene Expression
and Biochemistry
Rubisco; RbcS2 transcripts
PAL transcipts
Ace oxidase, catalase
Ascorbate peroxidase
Glutathione reductase
Phenolic glycosides
Tannins
Foliar nitrogen
C:N ratio of foliage
Starch
Gas Exchange
Amax lower canopy
A,^ whole canopy
Stomatal limitation
Stomatal conductance
Foliar respiration
Soil respiration
Microbial respiration
Stomatal density
Chlorophyll content
Chloroplast structure
O3flux
Growth and Productivity
Leaf thickness
Leaf size
Leaf area
CO2
d1
d
d
d
d
i
n.s.
d
i
d

n.s.
ii
d
d
n.s.
ii
ii
n.s.
d
i
d

i
i
i
03
d
i
i
n.s.
i
d
i
n.s.
n.s.
d

dd
dd
n.s.
di
i
d
n.s.
n.s.
d
d
ii

n.s.
d
d
CO2 + O3
dd
d
d
d
d
n.s.
i
d
ii
n.s.

id
n.s.
d
d
n.s.
n.s.
n.s.
n.s.
d
d
i

n.s.
d
n.s.
Reference
Wustman et al. (2001); Noormets et al. (200 la)
Wustmanetal. (2001)
Wustman etal. (2001)
Wustmanetal. (2001)
Wustmanetal. (2001)
Lindroth et al. (2001); Kopper and Lindroth
(2003a,b)
Lindroth et al. (2001); Kopper and Lindroth
(2003a,b)
Lindroth et al. (2001); Kopper and Lindroth
(2003a,b)
Lindroth etal. (2001)
Wustmanetal. (2001)

Takeuchi et al. (2001); Noormets et al. (200 la)
Noormets et al. (200 Ib); Sharma et al. (2003)
Noormets et al. (200 la)
Noormets et al. (200 la)
Takeuchi et al. (2001); Noormets et al. (200 Ib)
King etal. (2001)
Phillips et al. (2002)
Percy et al. (2002)
Wustmanetal. (2001)
Oksanen et al. (200 Ib); Takeuchi et al. (2001);
Wustmanetal. (2001)
Noormets et al. (200 la)

Oksanen etal. (200 Ib)
Wustmanetal. (2001)
Noormets etal. (200 Ib)
January 2005
9-274
DRAFT-DO NOT QUOTE OR CITE

-------
 1      2002).  This study also confirmed earlier observations of changes in trophic interactions
 2      involving keystone tree species, as well as important insect pests and their natural enemies
 3      (Table 9-26) (Percy et al., 2002; Holton et al., 2003; Awmack et al., 2003).
 4
 5      9.7.3  Landscape Condition
 6           In the SAB framework (Figure 9-19), landscape condition is assessed using the areal
 7      extent, composition of component landscape ecosystems or habitat types, and the pattern or
 8      structure of component ecosystems or habitat types (including biocorridors).  To date, no
 9      publications exist on the impacts of O3 exposure on landscape condition. The effects of O3
10      exposure have only been reported at the community or stand level (see Biotic Conditions,
11      below). The following is a description of current discussions by land stewards and of how
12      difficult it will be to quantitatively assess the effect of O3 exposure on landscape condition.
13           Landscapes are identified and preserved, such as National Parks, Class I Wilderness Areas,
14      etc., so that they are protected from the effects of O3 exposure by law.  Efforts to determine
15      whether landscapes have been affected by certain levels of exposure rely on valuation of
16      landscape and ecosystem components. Several different approaches of valuation have been
17      used, such as pathological (visible symptoms), biomass and allocation, and biogeochemical.
18           In the pathological approach, a "critical loads" concept is developed, with varying levels of
19      impact viewed as acceptable, interim targets, or as unacceptable. As an example, land managers
20      of Class I Wilderness Areas may consider a critical level acceptable if it resulted in no visible O3
21      symptoms to sensitive species. In concrete terms, sensitive species may respond to peak O3
22      exposures of 60 ppb (e.g., coneflower, Rudbeckia laciniata, in Great Smoky National Park;
23      Davison et al., 2003), and so the critical exposure level would be < 60 ppb for any hourly valued
24      during the growing season. An interim target would be that less than 5% of the sensitive plants
25      would have visible symptoms of < 15% of the leaf surface. An unacceptable level of O3
26      exposure would be any result more pronounced than the interim target.  The advantage of the
27      foliar injury approach is that large crews with relatively simple training can assess individual
28      species within the landscape and "see" the effect of the oxidant exposure.  There are several
29      disadvantages, however.  Some species (e.g., white fir) exhibit no foliar injury, but do have shifts
30      in biomass allocation in response to oxidant exposure (Retzlaff et al., 2001). Other species have
31      shown significant decreases in foliar injury due to needle loss, retranslocation of nutrients to

        January 2005                              9-275       DRAFT-DO NOT QUOTE OR CITE

-------
                       Sources
                        Above-ground Processes
                         (-) Rubisco activity
                         (+) Lipid Peroxidation
                         (-) Stomatal Control
                         (-) Phloem Loading
                         (+) Senescence
           Sinks
             Above-ground Processes
              (+) Antioxidant Synthesis
              (+) Repair Processes
              (+) Construction Costs
              (+) Tissue Respiration
              (-) Growth
              (±) Reproduction
                        Storage Carbohydrates
                         (-) Pool Size
                         (-) Concentration
                        Root Storage
                         (-) Pool Size
                         (-) Concentration
             Storage Carbohydrates
              (-) Pool Size
              (±) Concentration
                  Root Processes
                   (-) Growth
                   (±) Mycorrhizal
                      Colonization
                   (?) Turnover


                  Root Storage
                   (-) Pool Size
                   (-) Concentration
       Figure 9-19.  A conceptual diagram of processes and storage pools in sources and sinks
                     that are affected by ozone exposure. A plus (+) denotes an increase in
                     process rate or pool size, a minus (-) denotes a decrease in process rate or
                     pool size, and a plus-minus (+) denotes that both increases and decreases
                     have been reported in response to O3.  Primary effects in the shoots (1°) are
                     distinguished from secondary effects in roots (2°) since the primary site of
                     ozone action occurs in the leaves (from Andersen, 2003).
1      remaining foliage, with subsequent increased photosynthetic rate (Beyers et al., 1992).

2      In addition, the development of foliar symptoms within a species is related to sunlight and

3      microclimate (Davison et al., 2003).

4           In the biomass approach, O3 exposure resulting in a measureable decline in biomass

5      (usually of a target, sensitive species) is used to evaluate landscape condition.  The bulk of the

6      information available is from seedling responses to controlled chamber exposures, reviewed in

7      the previous section.  Some information exists for species in natural environments, but teasing

8      out concurrent stressors and finding adequate "controls" may be intractable.  For example, in a
       January 2005
9-276
DRAFT-DO NOT QUOTE OR CITE

-------
 1      long-term gradient of O3 exposure, N deposition, and drought, the site with the highest O3
 2      exposure had the greatest whole tree biomass (pole-sized trees) due to growth stimulation by N
 3      deposition (Grulke and Balduman, 1999).
 4           In the biogeochemical approach, changes in biogeochemical cycling are used to assess
 5      landscape condition. Species sensitive to O3 exposure have well known responses to O3
 6      exposure, including altered C allocation to below- and above-ground tissues, and altered rates of
 7      leaf production, turnover, and decomposition.  Changes in turnover rates of ephemeral tissues
 8      (leaves, fine roots) also affect nutritional status of the remaining tissue.  These shifts can affect
 9      overall carbon and N loss from the ecosystem in terms of respired C, and leached aqueous
10      dissolved organic and inorganic C and N. Instability in C and N pools and dynamics can affect
11      landscape-level nutrient dynamics even without significant inputs of N deposition. The endpoint
12      assessment is based on changes in water quality from or in the landscape, correlated to a defined
13      oxidant exposure level.  These approaches are linkable:  visible injury at a particular level could
14      be related to reduction in photosynthate, which would reduce whole plant biomass (and carbon
15      dynamics).  If O3 sensitive species are dominant within the landscape, then changes in C and N
16      dynamics over time would be expected to alter biogeochemical cycles.  Examples of forest types
17      that contain geographically extensive, O3-sensitive species that could be used in assessing
18      landscape-level changes include ponderosa pine in the western United States, tulip poplar
19      (Liriodendron tulipiferd) or loblolly pine (Pinus taeda L.) in the eastern U.S. decidous forests,
20      and Norway spruce (Picea abies (L.) Karst) in the Carpathian Mountains of eastern Europe.
21           Water quantity may also be  affected by O3 exposure at the landscape level. Moderately
22      high O3 exposure may affect the mechanism of stomatal opening (McAinsh et al., 2002),
23      resulting in sluggish stomatal opening and closing (Patterson and Rundel, 1989; Reich and
24      Lassoie, 1984).  During moderately high O3 exposure in a drought year, canopy transpiration was
25      greater for tulip poplar than on adjacent days with lower O3 exposure.  This could affect canopy
26      transpiration and landscape level water use (McLaughlin et al., 2004). Oxidant exposure (O3 and
27      NOX) may decrease the ability of exposed plants to close  stomata at night (Grulke et al., 2004),
28      thus increasing water loss from the landscape. Ecosystem models should aid in interpreting
29      O3-exposure effects at the landscape level.
30
        January 2005                             9-277       DRAFT-DO NOT QUOTE OR CITE

-------
 1      9.7.4  Biotic Condition
 2      9.7.4.1 Ecosystems and Communities
 3           The SAB framework described by Young and Sanzone (2000) (Figure 9-20) identifies
 4      community extent, community composition, trophic structure, community dynamics, and
 5      physical structure as EEA's for assessing ecosystem health.
 6
 7      COMMUNITY EXTENT
 8           Ecosystem function is dependent on areal extent, constituent species composition, trophic
 9      structure and its dynamics, and community physical structure. Genetic variation within species,
10      and the dynamics of the interactions that exist among different species and their biotic and
11      abiotic environment, are also involved (Agrawal and Agrawal, 2000). There are no reports of O3
12      exposure altering community distribution or extent.
13
14      COMMUNITY COMPOSITION
15           Significant changes in plant community composition resulting directly from O3 exposure
16      has been demonstrated in two forested areas: the mixed conifer forest of the San Bernardino
17      Mountains, CA and the mixed conifer forest of the Valley of Mexico near Mexico City. It is
18      also likely that community composition has changed in response to O3 exposure in the
19      coniferous forests of the Carpathian Mountains, but this has not yet been definitively shown.
20           The first forest communities shown to be affected by O3 were the Pinus ponderosa-
21      dominated stands of the San Bernardino Mountains in southern California (Miller, 1973). Miller
22      suggested that mixed forests of P. ponderosa, Pinus jeffreyi and Abies concolor were  changing
23      to predominantly A. concolor because of the greater sensitivity of the pines to O3. Significantly
24      greater mortality of young mature trees (50 to 99 years old) occurred in sites that also showed
25      higher foliar injury relative to sites that showed slight foliar injury (McBride and Laven, 1999).
26      For P. ponderosa, 33% of the trees in the high foliar injury sites died versus 7% of the trees in
27      the low foliar injury sites over the decade-long census.  In contrast, 24% of Abies concolor  died
28      in high foliar injury sites, whereas no trees died in slight injury sites. The authors suggested that
29      certain age classes were especially sensitive to O3 exposure because they are emerging into the
30      canopy, where higher O3 concentrations are encountered.  Future projections based on past
31      changes in community composition have been conducted for 2024 and 2074 (McBride and

        January 2005                              9-278       DRAFT-DO NOT QUOTE OR CITE

-------
 1      Laven 1999).  In their projections, the population of Pinusponderosa nearly disappears in all
 2      tree age classes, and the community is dominated by Quercus kelloggii in all tree age classes,
 3      followed by Calocedrus decurrem and Pinus lambertiana by the year 2074. Their projections
 4      do not account for potential changes in genetic structure of the more O3-sensitive species.
 5           In the Valley of Mexico, the closed forest structure changed to a woodland from high
 6      pollutant exposure (Miller et al., 2002). Cryptogamic community diversity also significantly
 7      declined in response to prolonged, extreme O3 exposure (Zambrano et al., 2002. Together,
 8      these two examples illustrate the potential for shifts in community composition in response to
 9      O3 stress.
10
11      TROPHIC STRUCTURE
12      A bove-Ground Structural
13           One of the first reports of trophic level interactions in natural communities was the
14      O3-induced predisposition of Pinus ponderosa to attack by bark beetles (Stark et al., 1968; Cobb
15      et al., 1968; Stark  and Cobb, 1969). Trees exposed to oxidant injury had lower resin production,
16      flow, and exudation pressure. Also, several attributes associated with tree defense against beetle
17      attack were compromised by oxidant exposure including sapwood and phloem moisture content
18      and phloem thickness (Pronos et al., 1999). Another trophic level has been implicated, in that O3
19      injured ponderosa  pine had the same rate of bark beetle infection, but healthy trees had greater
20      numbers of bark beetle predators and parasitoids (Dahlsten et al., 1997).  This suggests that O3
21      damage rendered the pines inhospitable for the natural enemies of the bark beetles.  Similar
22      findings were presented by Percy et al. (2002) for aphids whose abundance was increased in
23      young Populus tremuloides stands exposed to elevated O3. In that study, the levels of natural
24      enemies of alphids (ladybirds, lacewings, spiders and parasitoids) were significantly decreased
25      under elevated O3.
26
27      Below-Ground
28           Processing of plant-derived carbon compounds by soil  organisms comprising the soil food
29      web is a fundamental property of a functional and stable below-ground ecosystem (de Ruiter
30      et al., 1998; Wolters, 1998).  Soil food web organisms are responsible for recycling  nutrients and
31      for development of soil properties such as porosity, aggregate structure, water-holding capacity

        January 2005                              9-279       DRAFT-DO NOT QUOTE OR CITE

-------
 1      and cation exchange capacity.  A shift in food web species diversity or functional complexity in
 2      response to O3 stress may alter ecosystem processes.
 3           Evidence that soil organisms are affected by O3 indicates the potential for changes in soil
 4      food web structure and function. Since O3 does not penetrate the soil beyond a few centimeters,
 5      the proposed mechanism by which O3 alters soil biota is through a change in carbon input to
 6      soils (Andersen, 2003).  Ozone can alter C inputs to soil and hence soil processes through four
 7      different pathways: (1) leaf-litter quality and quantity (see Material Cycling, below), (2) carbon
 8      allocation to roots (see Physiological Status, below), (3) interactions among root symbionts, and
 9      (4) rhizodeposition.  The complex nature of the effects of O3 on trophic interactions and food
10      webs calls for additional basic research and modeling.
11           There have been no comprehensive studies on the effects of O3 on structural components of
12      soil food webs, however studies have shown that O3 affects free-living soil organisms of food
13      webs. In the few cases where soil microbial communities have been examined, O3 has led to
14      changes in bacterial and fungal biomass, and in some cases changes in soil enzyme activity.
15      Phillips et al. (2002) examined the effects of elevated CO2 and O3 on C flow through
16      heterotrophic microbial communities in soils collected from a FACE study in Wisconsin. Ozone
17      decreased abundance of fungal phospholipid fatty acids in aspen and birch-aspen plots but had
18      few other direct effects on measured soil parameters. The greatest effect of O3 was to eliminate
19      significant increases in microbial respiration resulting from elevated CO2, suggesting an
20      important role for O3 in altering C flow through soils.  Shafer (1988) found that O3 tended to
21      increase the number of fungal propagules and bacteria exhibiting phosphatase activity in the
22      rhizosphere of sorghum. Ozone in combination with simulated acid rain stimulated soil
23      arylsufatase activity (Reddy et al., 1991).  The response was observed at low concentrations, but
24      was reversed at high concentrations, suggesting a threshold level of O3, possibly involving
25      different mechanisms. Ozone significantly decreased soil microbial biomass in the fall after one
26      season of exposure in a wheat (Triticum aestivuni) and soybean (Glycine max) system (Islam
27      et al., 2000). Other studies have shown shifts in microbial and fungal biomass in response to O3
28      stress, but responses were variable (Scagel and Andersen, 1997; Yoshida et al., 2001).
29           Decreased allocation to roots associated with O3 exposure alters N fixation in legumes and
30      actinorrhizal species.  Ozone exposure was found to decrease nodulation in  a number of species
31      (Manning et al., 1971; Tingey and Blum, 1973). In alder (Alnus sermlata), host root cells of

        January 2005                              9-280        DRAFT-DO NOT QUOTE OR CITE

-------
 1      nodules showed cytoplasmic breakdown and lacked organelles when seedlings were exposed to
 2      27 d of O3 (Greitner and Winner, 1989).
 3           Ozone has been shown to affect mycorrhizal colonization (Ho and Trappe, 1981; McCool
 4      et al., 1982; Simmons and Kelly 1989; Adams and O'Neill, 1991; Edwards and Kelly 1992;
 5      Smith and Read, 1997).  Although short-term in nature, several studies have found enhanced
 6      mycorrhizal short-root formation under O3 stress. White pine (Pinus strobus) (Stroo et al.,
 7      1988), Norway spruce (Rantanen et al., 1994), Northern red oak (Quercus rubra) (Reich et al,
 8      1985), Douglas-fir (Pseudotsuga menziesii) (Gorissen et al., 1991), European silver-fir (Abies
 9      alba) (Wollmer and Kottke, 1990), and Scots pine (P. sylvestris) (Kasurinen et al., 1999) all
10      showed some increase in mycorrhizal presence when exposed to O3.  Others have shown
11      minimal or no effects of O3 on mycorrhizas (Mahoney et al., 1985; Meier et al., 1990;
12      Kainulainen et al., 2000). Stroo et al. (1988) found that percent infection increased from 0.02 to
13      0.06 ppm O3, then decreased from 0.06 to 0.14 ppm; the total number of short roots were
14      unaffected, however. In cases where stimulation was observed, the response was often noted
15      shortly after initiation of exposure, often at relatively low concentrations. Good examples of this
16      transitory response can be found in results with Norway spruce and Scots pine (Ratanen et al.,
17      1994; Kasurinen et al., 1999). In these studies, O3 increased mycorrhizal short roots initially but
18      differences were not evident by the end of the experiment.
19           There is evidence which suggests that decreased below-ground allocation associated with
20      O3 stress alters mycorrhizal host-symbiont compatibility. Edwards and Kelly (1992) found a
21      shift in fungal morphotypes present on loblolly pine roots, even though the number of
22      mycorrhizal short roots per gram fine root was not significantly affected by O3. Qiu et al. (1993)
23      found increased numbers of morphotypes present on O3-sensitive loblolly pine seedlings exposed
24      to O3.  Roth and Fahey (1998) found an interaction between O3 and acid precipitation treatments
25      on the composition of fungi forming ectomycorrhizae on red spruce saplings, possibly driven by
26      nutrient availability.  Carbohydrate requirements vary among fungal species (Bidartondo et al.,
27      2001), and O3 may affect species composition by altering carbohydrate availability in roots.
28      A shift in species dominance could lead to a change in successional patterns of mycorrhizal
29      communities.
30           In the few studies that examined root exudation in response to O3 exposure, O3 was found
31      to alter the quantity and quality of root exudates. McCool and Menge (1983) found a significant

        January 2005                              9-281        DRAFT-DO NOT QUOTE OR CITE

-------
 1      decrease in exudation of amino acids in tomato (L. esculentum) exposed to 300 ppb O3. McCrady
 2      and Andersen (2000) observed increased root exudation in nonmycorrhizal wheat (Triticum
 3      aestivum). No apparent change in root exudation was found in labeling studies of ECM
 4      ponderosa pine (Andersen and Rygiewicz, 1995). Inconsistency in the literature probably results
 5      from species differences and experimental protocols, however these examples illustrate the
 6      potential effects of O3 on rhizosphere carbon flux.
 7          Decreased carbon allocation to roots of O3-exposed plants may reduce root longevity and
 8      accelerate root turnover, increasing rhizodeposition of C and N. Fine root turnover decreased in
 9      mature northern red oak exposed to elevated O3 (seasonal exposure ranging from 152 to
10      189 ppm-h), whereas seedlings did not show any reduction in turnover (Kelting et al., 1995).
11      King et al. (2001) found a trend toward decreased live root biomass and increased dead root
12      biomass in aspen exposed to O3 in a FACE study, suggesting possible changes in both
13      product! on and 1 ongevity.
14          Other studies also suggest that O3 alters C flux to soils, resulting in changes in CO2 efflux
15      from soils. Both root respiration and soil CO2 efflux decreased from loblolly pine seedlings
16      exposed to O3 (Edwards, 1991).  Soil CO2 efflux increased in response to O3 in ponderosa pine
17      seedlings (Andersen and Scagel, 1997; Scagel and Andersen,  1997). No direct assessments of
18      hyphal growth and turnover in response to O3 stress have been conducted. Ozone decreased C
19      allocation to extrametrical hyphae of a ponderosa pine mycorrhiza, which might be expected to
20      decrease growth and increase hyphal turnover (Andersen and Rygiewicz, 1995).
21
22      COMMUNITY DYNAMICS, PHYSICAL STRUCTURE
23          One of the best-documented examples of change in long-term forest community dynamics
24      of dominant overstory trees occurred in the San Bernardino Mountains between 1968 and 1974
25      (reported in Miller,  1973; Miller et al., 1989). Plots were recently re-inventoried -25 years after
26      establishment (Arbaugh et al., 2003).  Of the six codominant canopy species, white fir showed
27      the greatest change, increasing in both numbers and bole growth for a 286% change in basal
28      area/ha in the San Bernardino Mountains.  Sugar pine basal area also increased significantly (by
29      334%), but this species represents only a small portion (1%) of the total basal area of the forest
30      sampled.  The most sensitive species (Miller et al., 1983), ponderosa and Jeffrey pine, had the
31      lowest increase in basal area/ha (76 and 62%, respectively).  These two species represented 72%

        January 2005                             9-282       DRAFT-DO NOT QUOTE OR CITE

-------
 1      of the basal area/ha of all stands inventoried.  Ponderosa pine had the greatest mortality rate of
 2      all canopy species inventoried (46%), followed by white fir and black oak (35% and 33%),
 3      Jeffrey pine (29%) and incense cedar and sugar pine (both 7%). In moist sites (at the western
 4      end of the San Bernardino Mountains), there was significant recruitment of incense cedar, white
 5      fir, and sugar pine.  Only one study directly attributed tree mortality to O3 exposure: it
 6      accounted for 7% of mortality in the Sierra and Sequoia National Forests (Carroll et al., 2003).
 7           Species diversity in the understory can be quite large, making studies of O3 effects on
 8      understory community dynamics very challenging. However, there have been some attempts to
 9      quantify understory responses, ranging from describing relative sensitivity to their visible
10      symptoms (Treshow and Stewart, 1973; Temple, 1999) to very  complex measures  of community
11      structure and composition (Westman, 1979, 1981). The lowest percentage cover and lowest
12      species diversity in  California coastal sage scrub was correlated with the highest O3 exposures as
13      estimated by extrapolation from the closest air monitoring stations (Westman  1979). The
14      understory also has  the potential to influence responses to O3 of dominant keystone species, as
15      has been shown in controlled experiments with bothponderosapine (Andersen et al., 2001) and
16      loblloly pines (Barbo et al., 2002). Barbo et al., (1998) exposed an early successional forest
17      community to ambient air, charcoal-filtered air, non-filtered air, and 2x ambient in the
18      Shenandoah National Park. They found changes in species performance, canopy structure,
19      species richness and diversity index consistent with the view that O3 can induce a shift in
20      vegetation dominance and community structure.
21           There have been few studies evaluating the effect of O3 exposure on the physical  structure
22      of natural ecosystems.  Despite an extensive array of allometric equations for conifers in the
23      west on United States (Ter-Mikaelian and Korzukhin, 1997), none appear to predict individual
24      tree shape in a site of moderate O3 exposure, suggesting that O3 may effect allometry (Grulke
25      et al., 2003a).  Canopy structural changes are also implied by the measure of canopy
26      transparency used in the FHM assessment. The loss of epiphytic lichens within the canopy is a
27      clear example of plant community structural change occurring along an O3 gradient (Nash and
28      Sigal, 1999; Zambrano et al., 2002).
29           As of yet, there have been no comprehensive studies on the effects of O3 on structural or
30      functional components below-ground (Andersen, 2003). Phillips et al. (2002) found evidence
31      for changes in the bacterial and fungal biomass below Populus  tremuloides and

        January 2005                              9-283        DRAFT-DO NOT QUOTE OR CITE

-------
 1      P. tremuloideslBetulapapyri/era stands exposed to elevated O3.  Subsequent study showed that
 2      O3 exposure decreased cellobiohydrolase activity in the soil microorganisms, driving the change
 3      in the microbial community (Larson et al., 2002).
 4
 5      9.7.4.2  Species and Populations
 6           Ozone can affect species and populations of species comprising ecosystems through
 7      changes in population size, genetic diversity, population structure and/or dynamics, and habitat
 8      suitability (Young and Sanzone, 2002). For example, if individuals of a species are lost due to
 9      O3 exposure, population size declines.  Often very young (e.g., conifer seedlings, see Section
10      9.7.4.3 below) and old individuals differ in their sensitivity, so that population structure also will
11      be altered by O3 exposure. If resource allocation to reproductive output is altered by O3
12      exposure, population dynamics will be altered. Communities dominated by O3 sensitive species
13      in the canopy or understory may be altered sufficiently for the habitat to become unsuitable for
14      other species.  Genetic selection acts on the individual plant, which represents a certain
15      proportion of the populations' genetic variation. If an O3  sensitive individual succumbs through
16      multiple stresses, including O3 stress, the genetic variation represented in  the population
17      generally declines, unless sensitive individuals have low inherent genetic  variability (e.g.,
18      Staszak et al., 2004).
19           While the concept of natural selection induced by O3 exposure and related changes in
20      natural plant communities has been around for a long time (Dunn, 1959; Miller et al., 1972), the
21      concept of evolution of O3 tolerance is still not widely accepted.  The unequivocal demonstration
22      that considerable genetic variation in O3 resistance exists within and between plant populations,
23      and that ambient levels of O3 may differentially affect fitness-related traits (i.e., growth, survival,
24      and fecundity), suggests that O3 may potentially drive the natural selection for resistant
25      individuals. Dunn (1959) presented circumstantial evidence that ambient O3 in the Los Angeles
26      area was high enough to drive the selection of O3 resistant Lupinus bicolor genotypes.  Since
27      Dunn's (1959) research on O3-induced population changes, researchers have demonstrated
28      differences in O3 tolerance among other plant populations. In the devastating forest decline
29      southeast of Mexico City, the remaining trees (primarily Abies religiosa in the "cemetery
30      forests") appear to be less affected by foliar injury than trees that were lost (Alvarado-Rosales
31      and Hernandez-Tejeda, 2002), despite continued, high O3 exposures (Bravo-Alvarez and

        January 2005                              9-284       DRAFT-DO NOT QUOTE OR CITE

-------
 1      Torres-Jardon, 2002). However, even the most convincing work in this field (Berrang et al.,
 2      1986, 1989, 1991), whhPopulus tremuloides, where a strong correlation between visible foliar
 3      injury after O3 exposure and maximum O3 concentration at the origin of the population was
 4      shown (Berrang et al., 1991), a change in gene frequency at any one site over time has not yet
 5      been demonstrated (Bell et al., 1991; Reiling and Davison, 1992b). Furthermore, the selection
 6      intensity of O3 has been questioned (Bell et al., 1991; Taylor and Pitelka, 1992; Taylor et al.,
 7      1994) and the emergence of O3 exposure since the 1950's as an environmental stressor may not
 8      have been long enough to affect tree populations with long generation times (Barrett and Bush,
 9      1991).
10           The loss of O3-sensitive individuals results in natural selection favoring O3-tolerant species
11      (Bradshaw and McNeilly, 1991). Increased levels of mortality of O3-sensitive individuals have
12      occurred for Pinus jeffreyi and P. ponderosa exposed to ambient O3 along the western slope of
13      the Sierra Nevadas (Peterson et al., 1987; Miller et al., 1998), for Pinus strobus exposed to
14      ambient O3 in southern Wisconsin (Karnosky, 1981), and for P. ponderosa in the San Bernardino
15      Mountains (Carroll et al., 2003).  In these examples, individuals that consistently had greater
16      foliar injury and lower needle retention were lost in repeated surveys. Ozone-induced loss of all
17      individuals except the most tolerant and breeding among the surviving individuals to yield more
18      more tolerant populations has not yet been demonstrated for plants exposed to O3, except for the
19      relatively short term (2 years) adaptation exhibited in Trifolium repens (Heagle et al., 1991) and
20      Plantago major (Davison and Reiling, 1995). Heagle et al. (1991) were able to show the
21      adaptation of a Trifolium repens population to elevated O3 in just two growing seasons.
22      Similarly, Davison and Reiling (1995) compared the O3 resistance of P. major populations
23      grown from seed collected from the same sites over a period of increasing O3.  The two
24      independent populations studied exhibited increased  O3 resistance, consistent with the idea of
25      selection for  O3 tolerance. Using random amplified polymorphic DNA primers, this team also
26      showed that the later populations are subsets of the earlier ones, consistent with in-situ evolution
27      rather than with catastrophic loss and replacement of the populations (Wolff et al., 2000). The
28      problem is determining whether spatial patterns in O3 resistance and changes in time are casually
29      related to O3, because there were very strong correlations with other factors (Reiling and
30      Davison, 1992a; Davison and Barnes, 1998). The potential for evolution of O3 resistance has
31      been clearly demonstrated by Whitfield et al. (1997)  in their study of O3 selection of common

        January 2005                             9-285       DRAFT-DO NOT QUOTE OR CITE

-------
 1      plantain (Plantago major L), where they showed that within a matter of a few generations, it
 2      was possible to increase O3 resistance in an initially O3-sensitive population. Wild radish
 3      (Raphanus sativus L.) developed O3 resistance after only one generation of exposure to O3
 4      (Gillespie and Winner, 1989).
 5           A third independent line of research suggesting O3 may be affecting the genetic diversity of
 6      wild plant populations was presented by Paludan-Miiller et al. (1999) who showed that northwest
 7      European provenances of European beech (Fagus sylvatica L.) were more sensitive to O3 than
 8      were southeast European provenances which had experienced higher O3 levels. Recent research
 9      on the genetic structure of 50-year-oldponderosapines in the San Bernardino Mountains
10      suggests that distinct differences in frequency of some alleles and genotypes occurred, with the
11      O3-tolerant trees being more heterozygous (Staszak et al., 2004). While both of these studies
12      were only correlational, they are consistent with previous studies of this type suggesting
13      O3-induced population changes.  Again, other environmental stressors besides O3 exposure could
14      have been involved in effecting change within these populations.
15           Natural selection for O3 tolerance can also be facilitated by reductions in fitness related to
16      lower seed yields of O3-sensitive species or individuals.  The impacts of O3 on reproductive
17      development, recently reviewed by Black et al. (2000) can occur by influencing:  (1) age of
18      flowering, particularly in long-lived trees that often have long juvenile periods of early growth
19      without flower and seed production; (2) flower bud initiation and development; (3) pollen
20      germination and pollen tube growth; and (4) seed, fruit, or cone yields and seed quality
21      (Table 9-24).  In addition, vegetatively propagated species can have lower numbers of
22      propagules under elevated O3 conditions (Table 9-24).
23           Several studies suggest that reproductive structures are clearly sensitive to O3 and that O3
24      can  affect fitness of plants by affecting either the sporophytic or gametic generations. Decreased
25      numbers of flower spikes and seed capsules per plant were found for  plantain growing under
26      elevated O3 (Reiling and Davison,  1992b; Pearson et al., 1996; Lyons and Barnes, 1998).
27      Similar responses  were seen for Brassica campestris L. plants exposed to a single dose of
28      100 ppb O3 for 6 h. Stewart et al. (1996) and Bosac et al. (1998) reported an increase of flower
29      bud abortion for oilseed rape (Brassica napus L.) similarly exposed to a short duration of
30      elevated O3.  Floral initiation period can be delayed in O3-sensitive plants, as was described for
31      dogbane (Apocynum androsaemifolium) grown under ambient O3 in the eastern United States.

        January 2005                              9-286        DRAFT-DO NOT QUOTE OR CITE

-------
 1      In one of the few comparisons of whole plant O3 sensitivities with that of male gametophytes,
 2      Hormaza et al. (1996) found a high correlation of relative O3 sensitivity of pollen tube elongation
 3      with that of O3 effects on net photosynthesis and relative growth rates for 6 species of fruit trees.
 4           Clearly, the concept of O3 -induced genetic change is an area that needs additional research
 5      attention. Repeated collections over time from wild populations receiving high O3 exposures to
 6      examine population responses and relative sensitivity changes, the sampling of genetic diversity
 7      along known O3 gradients, and the use of modern biotechnological approaches to characterize
 8      and quantify genetic diversity are useful approaches to test for O3-induced impacts on diversity
 9      in natural ecosystems.
10
11      9.7.4.3  Organism Condition
12      PHYSIOLOGICAL STATUS
13           The generalized effects of O3 exposure on plants are well known, and have been reviewed
14      from several viewpoints over the last decade (De Kok and Tausz, 2001; Heath and Taylor, 1997;
15      Pell et al., 1997; Schraudner et al.,  1997; Matyssek et al., 1995; Darrall  1989; Reich,  1987; U.S.
16      Environmental Protection Agency,  1996). The topic of individual species response and
17      modification of response by other factors has been addressed thoroughly in Sections 9.4.3 and
18      9.4.4 of this chapter. Here, physiological changes in response to O3 that have been hypothesized
19      to lead to changes in ecosystem structure or function are highlighted.
20
21      Above-Ground Responses
22           The first critical step leading to O3 response is uptake and movement of O3 by the leaves,
23      leading to changes in C and nutrient relations that are thought to alter plant growth and
24      competiveness (see Section 9.3). Ozone enters leaves through stomates, reacts with cell walls or
25      membranes, and starts a series of adverse reactions.  Cuticular uptake of O3 is believed to be
26      negligible (Kerstiens and Lendzian, 1989; Coe et al., 1995).  Once inside the leaf, O3  and its
27      byproducts lead to membrane disruption, chlorophyll breakdown, and decreased Rubisco levels
28      (Schweizer and Arndt, 1990). In turn, photosynthesis is decreased, as is stomatal conductance
29      (Weber et al., 1993). Also, O3 often leads to increased maintenance respiration, decreased foliar
30      nutrient content, and imbalances in tissue nutrient content and retention. When photosynthetic
31      pigments have been damaged, the pigment must be fully broken down (and/or new N and Mg

        January 2005                             9-287        DRAFT-DO NOT QUOTE OR CITE

-------
 1      must be taken up and transported to the leaf) for the pigment to be regenerated (Bjorkman and
 2      Demmig-Adams, 1995). Ozone exposure alters within-plant priorities for resources:  less C is
 3      available for allocation to roots and spring regrowth, and less foliar biomass is retained.  At the
 4      whole-organism level, O3 exposure decreases root mass (Grulke et al., 1998) and radial bole
 5      growth (Peterson et al., 1991; Muzika et al., 2002) with little impact on height growth. Visible
 6      symptoms of O3 injury vary between species and genotypes but often  include upper leaf surface
 7      stipple, chlorotic mottle, or large bifacial blotches of necrotic tissue. Premature senescence is
 8      typical of almost all O3-induced foliar damage.  All of these changes can alter the plant's ability
 9      to function in a broader ecosystem context.
10           The underlying mechanisms of O3-injury response in conifers, broadleaf deciduous trees,
11      and herbaceous species are assumed to be similar. However, several differences in long-lived
12      species are important at the ecosystem level.  Most of the research on  O3 effects has been
13      conducted on herbaceous species (i.e., crops).  Although a number of  native herbaceous species
14      have been identified as O3-sensitive. there are no published physiological studies on the effect of
15      O3 exposure on herbaceous or shrub species in situ.  In natural ecosystems, the majority of
16      species are not annuals, unless the  system is highly disturbed.  Nonetheless, response of crop
17      species to  elevated O3 may be used as an analog for native annual response: phenological
18      staging is accelerated (soybeans; Booker et al., 2004), thus "avoiding" additional O3 exposure.
19           Conifers have roughly half the stomatal conductance of deciduous broadleaf trees (Reich,
20      1987), leading to proportionally less O3 uptake at the same O3 exposure level.  Yet, except for
21      species of larch, individual conifer needles  are longer-lived and active over a greater portion of
22      the year.  Therefore, needle longevity can also work against the tree by increasing cumulative O3
23      exposure and exposure to other stressors. Increased needle longevity  is not always a
24      disadvantage, for example, conifers are physiologically active in early spring and late fall, during
25      times of lower oxidant concentrations.  These periods can contribute significantly to a net
26      positive annual C balance and from the standpoint of nutrient storage, are important in reparation
27      responses to pollutants. Patterson and Rundel (1995) reported that Jeffrey pine had significant
28      stomatal opening (one third that of a typical summer day) in mid-winter with snow on the
29      ground. At least pole-sized  and larger trees can mitigate reductions in C acquisition due to
30      oxidant exposure in the summer with C assimilation on favorable days in the winter.  The
31      interaction of environmental factors, plant phenology (the timing of growth events; birth and

        January 2005                              9-288       DRAFT-DO NOT QUOTE OR CITE

-------
 1      mortality of plant parts), physiological status (nutritional or moisture status; dormant or active
 2      growth within the year), and tree age (interannual differences in resource acquisition and
 3      allocation) all contribute to the complexity of long-lived species (and hence ecosystem) response
 4      to O3 exposure.
 5           One widely observed response to O3 exposure is premature leaf loss. As noted above,
 6      premature leaf loss may reduce O3 uptake during high-O3 years, but it has several negative
 7      consequences.  Early leaf loss results in reduced C uptake through photosynthesis.  Premature
 8      needle loss  also results in less N retranslocation compared to normally senescing leaves, which
 9      reduces whole plant N balance (Fenn and Dunn, 1989) and carbohydrate availability for
10      overwinter storage  (Grulke et al., 2001).  Because of such effects  accumulated over several years
11      of O3 exposure,  subsequent-year C and N reserves can be affected (Andersen et al., 1991).
12           Conversely, a series of drought years can decrease O3 uptake, as well as reduce C and
13      nutrient acquisition, altering resource allocation to defenses (e.g., antioxidants) (Grulke et al.,
14      2003b) (or resins) against insect infestation, rendering the tree more susceptible to O3 injury.
15      Conifers have thicker cuticles than either broadleaf deciduous or herbaceous species. Continued
16      O3 exposure may compromise cuticular integrity  (Percy et al.,  1994). Once cuticular integrity is
17      breached, individual leaves (needles) are likely to be excised, thus contributing generally to
18      defoliation and reduced C acquisition.
19           With the exception of the extensive research conducted on mature tree response to O3
20      exposure in California forests (Peterson et al., 1987, 1991, 1995, Arbaugh et al., 1998; Grulke
21      1999; Grulke and Balduman 1999; Grulke et al.,  1996, 1998, 2001, 2002b, 2003a,b, 2004;
22      Weiser et al., 2002), the vast majority of studies of O3 effects on forest trees have been
23      conducted on young seedlings (Chappelka and Samuelson, 1998) and little is known about
24      acclimation to O3 (Skarby et al., 1998). Chamber exposure studies can be used to document
25      foliar symptoms and develop response variables for the whole plant.  These response variables
26      can then be field tested  on mature trees using correlative analyses (e.g., Grulke and Lee, 1997;
27      Grulke et al., 2003b). Without the initial  work in chamber exposure studies, field responses to
28      O3 exposure would be difficult to verify and distinguish from other concurrent stressors.
29           Predicting mature tree responses to  O3 solely from seedling  response studies is complex,
30      because seedlings or saplings do not necessarily respond to O3 in the same way as mature trees
31      (Norby et al., 1999; Karnosky, 2003). Ozone has been found to have stronger effects on leaf

        January 2005                              9-289        DRAFT-DO NOT QUOTE OR CITE

-------
 1      function in younger rather than older trees (Kolb and Matyssek, 2001). Each component
 2      physiological attribute "matures" at a different rate.  Gas exchange patterns differ between
 3      seedlings and mature trees. For example, leaf respiration of juvenile ponderosa pine was greater
 4      than that of mature trees (Momen et al., 1996). In the conifers tested, the highest gas exchange
 5      rates (and by inference stomatal uptake of O3) are found in seedlings (e.g., in scions of red
 6      spruce, Rebbeck and Jensen,  1993; giant sequoia, Grulke et al., 1994; ponderosa pine, Grulke
 7      and Retzlaff, 2001; and Norway spruce, Wieser et al., 2002). Patterns of biomass (Grulke and
 8      Balduman, 1999) and carbohydrate allocation (Grulke et al., 2001) differs between immature and
 9      mature trees.  Pole-sized trees had greater reduction in root, foliar, and bole carbohydrate
10      concentrations than did old growth trees. Antioxidant defenses vary with both tree age and
11      needle age (Tegisher et al., 2002). Based on all attributes measured in both ponderosa pine and
12      giant sequoia, the youngest tree age considered representative of mature trees was 20 years old
13      (Grulke et al., 1996; Grulke and Retzlaff, 2001).  In some broadleaf deciduous tree species,
14      seedlings are more conservative, and mature trees have greater gas exchange rates, as is the case
15      for Quercus robur (Edwards et al., 1994; Kelting et  al., 1995; Samuelson and Kelly, 1996, 1997)
16      and Fagus sylvatica (Braun et al., 1999). In another broadleaf deciduous tree species (cherry,
17      Prunus serotina), gas-exchange rates of seedlings were faster, but total O3 flux to leaves of
18      seedlings was lower than that of mature trees due to differences in leaf ontogeny (Frederickson
19      etal., 1995, 1996).
20           Nitrogen deposition modifies the effects of oxidant exposure through several  offsetting
21      physiological mechanisms (see Section 9.4.4). Nitrogen deposition,  in wet or dry particulate
22      form, ultimately increases site fertility, but increased soil N availability decreases C allocation to
23      roots, further exacerbating the effects of O3 exposure on roots (Grulke et al., 1998). Increased N
24      availability also increases foliage turnover: fewer needle age classes are retained (Gower et al.,
25      1993). Therefore, the combination of both increased N and O3 exposure increases foliar
26      turnover. Finally, N deposition and increased plant N nutrition can increase stomatal
27      conductance, leading to increased O3 uptake.  Alternatively, increased N counteracts the effect of
28      O3 on photosynthesis by increasing photosynthetic pigments and enzymes. Nitrogen deposition
29      may mitigate the degree of foliar injury from oxidant pollution via higher available N for
30      reparation of photosynthetic pigments.  Nitrogen amendments also modify the antioxidant
31      defense system in  complex ways (Polle, 1998).

        January 2005                               9-290        DRAFT-DO NOT QUOTE OR CITE

-------
 1           Attributes of O3 injury to trees (foliar injury, needle retention, and canopy transparency), as
 2      well as presence of pathogens and insect infestation, are routinely inventoried in established
 3      sample plots distributed on Federal lands across the United States (Forest Health Protection,
 4      USDA Forest Service).  Foliar injury to several widespread, herbaceous species nationally
 5      recognized as sensitive (bioindicators) is also assessed (NFS, 2003). This assessment is part of a
 6      larger assessment of forest tree growth and dynamics (the Forest Inventory and Analysis
 7      Program; Smith, 2002; Smith et al., 2003). Risk of O3 injury is then estimated for the dominant
 8      forest tree species in the sample plots.  For example, 12% of sampled black cherry (Primus
 9      serotina), 15% of loblolly pine (Pinus taeda), and 24% of sweetgum (Liquidambar styraciflud)
10      were found to be in the highest risk category in the northeast and mid-Atlantic states (Coulston
11      et al., 2003). In the Carpathian Mountains, 12 to  13% of all trees (broadleaf and coniferous)
12      have greater than 26% crown defoliation (Badea et al., 2002). In general, broadleaves (primarily
13      beech) trees were less affected (8 to 45%) than spruce (up to 37%) and fir (up to 50%)
14      (Grodzinska et al., 2002).  Ozone injury was directly correlated with cumulative O3 exposure in
15      the Sierra Nevada Mountains (Arbaugh et al., 1998); with the best correlation being found across
16      sites where > 90% of the trees had O3 injury. Although direct links of visible foliar symptoms
17      induced by O3 to adverse effects on biomass are not always found, visible foliar symptoms have
18      been linked to decreased vegetative growth (Peterson  et al., 1987; Karnosky et al., 1996; Somers
19      et al.,  1998), as well as reproductive function (Black et al., 2000; Chappelka, 2002).
20           Foliar O3 injury has also been associated with adverse effects on competitive ability and
21      survival in forest communities (Karnosky, 1981; McDonald et al., 2002; Karnosky et al., 2003b).
22      Competition can alter organism condition and affect susceptibility to O3. Ponderosa pine
23      seedlings were more susceptible to O3, as determined by decreased plant biomass, when grown
24      in competition with blue wild-rye grass (Andersen et al. 2001).  Similarly, the magnitude of O3
25      effects on height and diameter growth depended on the competitive status ofPopulus
26      tremuloides trees (McDonald et al., 2002). These studies show the importance of including
27      competition as a concurrent stressor in assessing whole plant responses to O3.  The age of the
28      community ("time since disturbance") may also affect the ability of individuals to effectively
29      respond to O3.  Unfortunately, the vast majority of O3  studies have been conducted on
30      open-grown plants, often grown in pots where competition is absent both above and below
31      ground.

        January 2005                              9-291       DRAFT-DO NOT  QUOTE OR CITE

-------
 1           Clearly, age-dependent O3 responsiveness and juvenile-mature correlations remain
 2      important research questions in attempting to scale up to ecosystem level responses. Patterns of
 3      allocation between root, stem, and leaf differ between immature and mature trees. Tree
 4      architecture varies with tree age, and leaf area distribution in space and time may change in
 5      response to elevated O3. All of these factors influence gas exchange in the canopy.
 6      Furthermore, there may be few generalities that can be made from seedling to mature tree
 7      response to O3 within a plant functional group (Norby et al., 1999; Karnosky, 2003).
 8      Consequently, modeling ecosystem response is limited to either dealing with mono-specific
 9      plantations or assigning average responses to a mix of species.
10
11      Below-Ground Responses
12           The effect of O3 on the soil ecosystem is thought to occur through physiological changes in
13      the root and interactions with soil organisms (Andersen, 2003).  Comparatively little is known
14      about how changes in root growth and metabolism are translated through the soil food web,
15      resulting in changes in soil and hence,  ecosystem processes.  An overview of physiological
16      changes likely to lead to changes at the ecosystem level is provided below.
17           Ozone stress decreases carbon allocation to roots (Manning et al.,  1971; McLaughlin and
18      McConathy,  1983; McCool and Menge, 1983; Cooley and Manning, 1987; Gorissen and van
19      Veen, 1988; Spence et al., 1990; Gorissen et al.,  1994; Rennenberg et al., 1996; U.S.
20      Environmental Protection Agency, 1996). Since roots are often dependent on current
21      photosynthate for their structural development (van den Driessche, 1978; Ritchie and Dunlap,
22      1980; Marshall and Waring, 1985; van den Driessche, 1991), C-limiting stresses such as O3 can
23      have rapid and significant effects on root growth. In many cases, decreased allocation to roots in
24      response to O3 occurs quickly, with reductions in root growth occurring within one growing
25      season (Gorissen and van Veen, 1988;  Spence et al.,  1990; Gorissen et al., 1991; Andersen and
26      Rygiewicz, 1991, 1995; U.S. Environmental Protection Agency, 1996).  Decreased C allocation
27      below ground is often associated with  decreased root-shoot ratio, but observed responses in
28      root-shoot ratio are highly variable owing to several factors including intra- and interspecies
29      variation, culture conditions, and ontogenetic drift (Reich, 2002).  Root-shoot ratio is a
30      point-in-time measurement that does not include C lost to exudation, respiration or turnover.
        January 2005                              9-292       DRAFT-DO NOT QUOTE OR CITE

-------
 1      Therefore, biomass and ratios of biomass (such as root-shoot ratio) do not necessarily reveal
 2      physiological changes in response to O3 stress.
 3           Decreased C acquisition leads to reduced carbohydrate levels and storage pools in
 4      O3-exposed plants (Tingey et al., 1976; Ito et al., 1985; Cooley and Manning, 1987; Rebbeck
 5      et al., 1988; Gorissen et al., 1994; Andersen et al., 1997; McLaughlin et al., 1982). Although it
 6      is difficult to quantify changes in the field, Grulke et al. (1998) found decreased medium and
 7      fine root biomass with increased pollutant load across an O3 gradient in southern California.
 8      Coarse and fine root starch concentrations also were lowest in mature trees at the most polluted
 9      site (Grulke et al., 2001).  The effects of O3 could not be completely separated from other known
10      stresses across the pollutant gradient, but  it appeared that O3 was an important factor in the
11      patterns observed.
12           Decreased storage pools can lead to carry-over effects on root growth that are compounded
13      over time. Decreased carbohydrate storage pools were associated with decreased root growth
14      during the spring following exposure to O3, even in the absence of additional O3 exposure
15      (Andersen et al., 1991, 1997).  Decreased spring root growth was attributed to decreased stored
16      C reserves as well as to premature loss of older foliage age classes during the previous fall.
17      Aside from the loss of photosynthetic surface area associated with premature senescence, early
18      loss of foliage in the fall occurs when allocation to roots is at a maximum in many species
19      (Kozlowski and Pallardy, 1997). Older needle age classes preferentially allocate photosynthate
20      basipetally to stems and roots (Rangnekar et al., 1969; Gordon and Larson, 1970), and the loss of
21      older needles in the fall during allocation to root growth and storage, and in the spring  during
22      periods of root growth, preferentially impacts roots and root processes.
23           Ozone has also been shown to affect root metabolism as evidenced by changes in root
24      respiration. Edwards (1991) found decreased root and soil CO2 efflux during a 2-year exposure
25      of loblolly pine to O3. Fine root respiration increased in mature red oak exposed to O3, while
26      total  soil CO2 efflux increased in the spring and decreased in the summer and fall (Kelting et al.,
27      1995).  The authors attributed increased root respiration to increased nutrient uptake  in support
28      of increased demands in the shoot. Ozone decreased root system respiration in aspen after 12
29      weeks of exposure, but the decrease was closely associated with decreased root biomass and
30      probably  not metabolic processes (Coleman et al., 1996).  Whether other metabolic shifts occur
31      in the roots of plants exposed to O3 needs to be examined.

        January 2005                             9-293        DRAFT-DO NOT QUOTE OR CITE

-------
 1           Measurable effects on roots may occur before effects on shoots are observed because
 2      shoots have immediate access to C for repair and compensation whereas roots must compete
 3      with shoots for C.  Mortensen (1998) found decreased root but not shoot growth in Betula
 4      pubescens at O3 exposures of 42 nMol mol-L (applied 12 h d"1), whereas both root and shoot
 5      growth were reduced at higher exposures.  Chromosomal aberrations were found in root tips of
 6      Norway spruce exposed to O3, even in the absence of biochemical changes in needles (Wonisch
 7      et al., 1998, 1999).  Using relatively high O3 concentrations (0.15 ppm O3 6 h d"1), Hofstra et al.
 8      (1981) found metabolic changes in Phaseolus vulgaris root tips prior to the development of leaf
 9      injury.  Morphological changes in root tips occurred within 2 to 3 days, and metabolism declined
10      within 4 to 5 days of initiation of O3 exposure.
11           Feedback signals from roots can influence the degree of O3 response. Stolzy et al. (1964)
12      exposed tomato roots (Lycopersicon esculentum) to periods of anerobic conditions and followed
13      a change in leaf susceptibility to O3. An exposure of roots to low oxygen conditions for 3 h did
14      not alter photosynthesis, but foliar damage was decreased when the roots were subsequently
15      exposed to O3. In this case, a signal originating in the root appeared to alter leaf sensitivity to
16      O3, the  signal possibly being hydraulic in nature and leading to decreased O3 uptake.
17
18      SYMPTOMS OF DISEASE OR TRAUMA,  SIGNS OF DISEASE
19           Although insects and diseases are dynamic components of forest ecosystems, trees can be
20      especially susceptible to outbreaks due to the presence of multiple stressors such as drought and
21      pollutant exposure. Ozone  can have direct effects on insect or disease organisms, indirect effects
22      on the insect or pathogen through changes to the host, and direct or indirect effects on natural
23      enemies of the insect or pathogen (Pronos et al., 1999). A full discussion of O3 effects on insect
24      and pathogen interactions can be found in Section 9.4.
25           Although the multitude of interacting factors makes it difficult to identify causative factors
26      in the field, some recent examples suggest a role for O3 in the timing or magnitude of disease
27      attacks  in the field.  After periods of drought stress (such as 1995 in central Europe), the
28      incidence of bark beetle (Ips spp.) appears to increase. In 1998 and  1999, the mean daily capture
29      of Ips was lowest in plots with low O3 exposure; the converse was also true (Grodzki et al.,
30      2002).  Elevation confounded the relationships, but the differences in Ips frequency in relation to
31      O3 concentrations were highly significant at lower elevations.  In the Valley of Mexico, a 1982

        January 2005                             9-294       DRAFT-DO NOT QUOTE OR CITE

-------
 1      to 1983 drough was documented, but not described as precipitating a bark beetle attack in the
 2      early 1980's. However, a link between air pollutant exposure and bark beetle attacks were
 3      implicated, because attacked trees were already O3-stressed at the time of the bark beetle attack
 4      (Alvarado-Rosales and Hernandez-Tejeda, 2002).
 5           An early study showed that oxidant exposure predisposedponderosapine to the root
 6      pathogen Fames annosus (James et al., 1980). Both root diseases (Pronos et al., 1999) and O3
 7      exposure (Grulke and Balduman, 1999) can each reduce root biomass, leading to increased
 8      drought stress, insect attack, and subsequent windthrow or death. Trees may take several years
 9      to die, and the patterns of precipitation and annual total rainfall interact to drive the level of
10      drought stress experienced by the tree (Pronos et al., 1999). Additional research is necessary to
11      fully understand the complex interactions occurring between O3 stress and other biotic stresses.
12
13      9.7.5 Ecosystem, Chemical, and Physical Characteristics (water, soil)
14      9.7.5.1 Nutrient Concentrations, Trace Inorganic and Organic Chemicals
15           Ozone exposure reduces the nutritional content of tissues, as well as causing elemental
16      imbalances. Foliar nutrient content may be too high (toxic) or too low (deficient), but the
17      relative amounts and ratios among all nutrients can also result in imbalances.
18           Although N deposition and foliar N content increased with O3 exposure in the San
19      Bernardino Mountains, K, Mg, Fe, and Al were all also higher in ponderosa pine at sites more
20      exposed to air pollution (Poth and Fenn, 1998). Trees with greater foliar injury (due to O3
21      exposure) had higher current year needle concentrations of P, K, Zn, and Fe than trees  at the
22      same site that were less injured. In drought-stressed ponderosa pine with O3 exposure, foliar N
23      was also elevated and retained in the remaining needles (Temple et al., 1992). At a relatively
24      clean site in the eastern San Bernardino Mountains, N, P, and K were efficiently readsorbed, but
25      P remaining in the foliage was relatively high compared to defined thresholds. The fact that
26      other elements were modified besides the N being deposited emphasizes the degree of chemical
27      imbalance in the tissue. Foliar micronutrients were within the normal ranges reported for
28      ponderosa pine (Poth and Fenn, 1998; Powers, 1981). Because both N deposition and O3
29      exposure reduce root biomass, it was unlikely that the foliar nutrient content was higher due to
30      greater uptake. Instead, it appaears that retranslocation from senescing tissue was responsible.
31      Across a pollution gradient in the Carpathian Mountains in eastern Europe, only S/N (expected

        January 2005                              9-295       DRAFT-DO NOT QUOTE OR CITE

-------
 1      due to high S deposition) and Fe/Mn ratios were out of balance relative to established norms.
 2      The S was relatively high due to SO2 deposition, and the Fe was relatively high due to smelter
 3      plumes.  No imbalances could be directly attributed to O3 exposure (Mankovska et al., 2004).
 4
 5      9.7.6 Ecological Processes
 6      9.7.6.1  Energy Flow
 7           All green plants generate and use energy-containing C compounds through the processes of
 8      photosynthesis and respiration.  Whole-plant C uptake is dependent on photosynthesis rates, leaf
 9      area, and leaf phenology. The effects of O3 at the site of action in the leaf are discussed in
10      Section 9.3. Here, the main focus is on whole-plant carbon dynamics resulting from changes in
11      C acquisition or use under O3 stress.
12           In natural ecosystems, O3 has been shown to depress photosynthesis in sensitive tree
13      species including Pinusponderosa (Miller et al., 1969; Weber et al., 1993; Takemoto et al.,
14      1997; Grulke et al., 2002b) andPopulus  tremuloides (Coleman et al., 1995a; Yun and Laurence,
15      1999; Noormets et al., 2001a,b; Sharma et al., 2003). In a study of mature Jeffrey pine, trees in
16      mesic microsites had greater O3 uptake over the growing season in comparison to  trees in xeric
17      microsites (Grulke et al., 2003a) and greater O3 uptake was correlated with lower mid canopy
18      needle retention, lower branch diameters, and lower foliar N content (Grulke et al., 2003a).
19      Chamber studies have also shown negative effects of O3 on tree seedling canopy structure
20      (Dickson et al., 2001) and leaf area (Neufeld et al., 1995;  Wiltshire et al., 1994). It is well
21      known from chamber and field studies that O3 exposure is correlated with lower foliar retention
22      (Miller et al., 1963; 1972; Karnosky et al.,  1996; Grulke and Lee, 1997; Pell et al., 1999; Topa
23      etal., 2001).
24           In contrast to the relatively consistent findings for photosynthesis, O3 effects on respiration
25      have been  more variable. Stem respiration was unaffected by O3 exposure (Matyssek et al.,
26      2002), suggesting that construction costs of new stems are not affected by O3.  However, the bole
27      represents  a relatively large storage pool of carbohydrates in mature trees, and the timing of
28      phenological events among individual trees may help to confound the ability to statistically
29      detect differences in stem respiration across pollutant gradients (Grulke et al., 2001).  Below-
30      ground respiration has been found to both increase and decrease in response to O3, depending on
31      the approach and timing of CO2 measures (Coleman et al., 1996;  Andersen and Scagel, 1997;

        January 2005                              9-296       DRAFT-DO NOT QUOTE OR CITE

-------
 1      Scagel and Andersen, 1997; King et al., 2001). The decreased soil respiration is thought to be
 2      due to reduced root growth under O3 exposure, but could also be partially explained by
 3      decreased microbial respiration in response to O3. Additional research is necessary to identify
 4      the role of O3 in affecting root vs. heterotrophic respiration, particularly over long time intervals.
 5           Carbohydrate availability and use influence the degree to which plants respond to O3
 6      exposure. A model simulation of the effect of O3 exposure on bole growth of Pinusponderosa
 1      showed a 15% reduction in mass (Weber and Grulke, 1995), largely influenced by differences in
 8      carbohydrates allocated and partitioned in repair  processes elsewhere in the tree. Foliar
 9      respiration is thought to increase under elevated O3 as maintenance costs (energy needs) of
10      leaves damaged by O3 are higher than normal (Grulke and Balduman, 1999; Noormets et al.,
11      200Ib), but differences in foliar respiration are subtle and difficult to detect statistically.  Foliar
12      carbohydrate studies also suggest that more C is used under O3 stress for repair processes (Topa
13      et al., 2001; Grulke et al., 2001) which would result in increased respiration. Ozone exposure
14      also reduced enzymatic activities  of carbohydrate metabolism related to the breakdown of
15      sucrose (Einig et al., 1997).  Changes in soil respiration in response to O3, even though O3 does
16      not  penetrate into the soil, illustrates the tight coupling of plant C balance and soil biota and
17      illustrates the potential role O3 plays in altering ecosystem C balances (Andersen,  2000).
18           Ozone can affect plant allometry through changes in energy use, potentially  affecting net
19      primary production (NPP) at larger scales. The net effect of O3 impacts on photosynthesis and
20      respiration for sensitive components of natural ecosystems is that height growth (Isebrands et al.,
21      2001;  Oksanen, 2003b); and radial growth (Peterson et al., 1987,  1991; Isebrands et al., 2001;
22      Terrazas and Bernal-Salazar, 2002; Oksanen, 2003b) can be negatively affected by O3. This has
23      been extrapolated to decreased NPP (Hogsett et al., 1997; Laurence et al., 2000).
24           Energy flow in plant communities can be altered by O3 through changes in C allocation.
25      It is well known that elevated O3 affects C allocation to roots (Coleman et al., 1995b;  Andersen
26      et al.,  1997; Grulke et al., 1998; Grulke and Balduman,  1999; Grulke et al., 2001)  by decreasing
27      or inhibiting phloem loading of carbohydrates (Grantz and Farrar, 1999; Landolt et al., 1997), or
28      of carbohydrate metabolism (Einig et al., 1997).  This leads to depressed root growth  (Andersen
29      et al.,  1991; Coleman et al.,  1996; Grulke et al., 1998) and the potential for plant communities to
30      have an increased susceptibility to drought through altered root-shoot balance. Furthermore, it
31      can negatively affect below ground food webs (Scagel and Andersen, 1997; Phillips et al., 2002).

        January 2005                              9-297        DRAFT-DO NOT QUOTE OR CITE

-------
 1           Another energetically costly response to O3 exposure is that the production of defense
 2      compounds, such as antioxidants, tend to increase under elevated O3 conditions (Sheng et al.,
 3      1997; Tausz et al., 1999b; 2002). Antioxidants help the plant scavenge free radicals before they
 4      can cause damage to membranes or cell walls, but they demand C for production such that
 5      growth can be adversely affected. In mature Jeffrey pine, stomatal uptake of O3 elicited one
 6      complex of antioxidant defenses in mesic microsites, while endogenously generated free radicals
 7      in the chloroplast elicited a second complex of antioxidant defenses in xeric microsites (Grulke
 8      et al., 2003b).
 9
10      9.7.6.2 Material Flow
11           Plants as producers are responsible for using inorganic atmospheric C and reducing it into
12      organic forms used by consumers, thus driving nutrient processes in ecosystems. Ozone has the
13      potential to disrupt material flow through organic C cycling and changes in nutrient cycling,
14      particularly N and P cycling. Although there is indirect evidence that O3 is disrupting C and
15      nutrient cycling at the ecosystem level, there is little direct evidence that O3 alters nutrient
16      processing at ecosystem scales.
17           The greatest annual nutrient and C input to ecosystems is from foliar and root turnover.
18      Excision of plant parts and whole plant mortality are potentially much larger, but syncopated,
19      ecosystem inputs.  Ozone exposure alters C cycling in the ecosystem by affecting the
20      within-plant C allocation and partitioning of dominant, O3-sensitive plants, and through chemical
21      composition and rate of decomposition of sloughed plant parts (roots, branches, leaves)
22      (Figure 9-19).
23           In addition to O3-induced changes in the quantity of C and nutrient inputs into ecosystems,
24      O3 also can alter the nutrient quality of inputs.  Ozone exposure alters nutrient levels in the
25      foliage (Boerner and Rebbeck,  1995; Lindroth et al., 2001; Fenn and Poth, 1998; Momen et al.,
26      2002) and affects the C:N ratio (Andersen et al., 2001; Lindroth et al., 2001; Grulke and Lee,
27      1997; Grulke et al., 2003b). Concentrations of compounds such as tannins, lignin and phenolics
28      (Kim et al., 1998; Findlay et al., 1996; Baumgarten et al., 2000; Saleem et al., 2001) are also
29      affected by O3  exposure, which in turn alters decomposability (Fenn and Dunn, 1989) and litter
30      buildup in the ecosystem.
        January 2005                             9-298        DRAFT-DO NOT QUOTE OR CITE

-------
 1           There are several possible pathways by which O3 may affect litter quality and, hence, litter
 2      decomposition, thus altering nutrient flow in ecosystems. These include altered C quality,
 3      altered nutrient quality, and alteration of leaf surface organisms important in decomposition
 4      pathways.  For example, yellow poplar (Liriodendron tulipiferd) and black cherry (Prunus
 5      serotina) litter exposed to O3 showed greater N loss during decomposition than charcoal filtered
 6      controls, although mass loss did not vary among O3 treatments (Boerner and Rebbeck, 1995).
 7      Subsequent studies  showed that although foliar N was not affected by O3 exposure in yellow
 8      poplar leaves, foliage decomposed more slowly (Scherzer et al., 1998).  Other studies have also
 9      shown a change in foliar N concentration in response to O3 treatment, affecting the C/N ratio and
10      possible litter quality (Andersen et al., 2001).
11           In some cases, it appears that N remobilization from foliage into the plant is not complete
12      at the time of foliage abscission in O3-exposed plants (Findlay and Jones, 1990; Stow et al.,
13      1992; Matyssek et al., 1993; Patterson and Rundel, 1995). Greater N content of senesced litter
14      could increase rates of decomposition. When O3-exposed cottonwood (Populus deltoides) leaves
15      abscissed at the same time as control leaves, they decomposed at similar rates; however,
16      prematurely senesced foliage from O3-exposed  cottonwood decomposed more slowly than
17      controls despite their higher N content (Findlay and Jones, 1990; Findlay et al., 1991). Higher N
18      in senesced leaves appeared to be related to organic complexes formed by bound phenolics in
19      O3-exposed leaves,  that made the litter less palatable to decomposers, thereby slowing
20      decomposition rates (Jones et al.,  1994; Findlay et al., 1996).  Increased phenolics also have
21      been found in European silver birch (Betulapenduld) exposed to O3 (Saleem et al., 2001).
22           Carbon quality in leaf litter also changes in O3-exposed foliage. Compositional changes in
23      leaf structural characteristics, such as lignin content, would be expected to alter rates of litter
24      decomposition (Fogel and Cromak,  1977; Meentemeyer,  1978; Kim et al., 1998).  Blackberry
25      (Rubus cuneifolus) litter exposed to  elevated O3 had greater permanganate lignin than control
26      treatments, which was inversely related to mass-loss rates in decomposition studies (Kim et al.,
27      1998).
28           Ozone may affect early stages of decomposition by altering populations of leaf surface
29      organisms before or after senescence. Magan et al. (1995) found a shift in phyllosphere fungi on
30      Scots pine (Pinus sylvestris\ Sitka spruce (Picea sitchensis\ and Norway spruce (Picea abies)
31      exposed to O3, but the potential effect of these changes on subsequent litter decomposition was

        January 2005                              9-299       DRAFT-DO NOT QUOTE OR CITE

-------
 1      uncertain.  The slowest decomposition rates of pre-exposed blackberry leaves were found when
 2      senesced foliage was exposed to O3 during decomposition, suggesting a possible direct effect of
 3      O3 on microorganisms in decomposing litter (Kim et al., 1998). Whether O3 concentrations at
 4      the soil surface influence initial stages of litter decomposition remains to be addressed.
 5           Ozone exposure also reduces nutritional content of foliage because of the degradation of
 6      chlorophyll. Reconstruction of chlorophyll may be limited by nutritionally poor soils or low soil
 7      moisture, as well as alteration of root uptake by O3 exposure and other stressors. Foliar exposure
 8      to O3 may also increase leaching of nutrients (Kerstiens and Lendzian, 1989).  Ozone exposure
 9      promotes early senescence of foliage (Miller and Elderman, 1977; Heath and Taylor,  1997), with
10      higher nutrient content than if excised later in the growing season (Poth and Fenn, 1998).
11           Since O3 can slow decomposition through changes in leaf quality and quantity, leaf litter
12      can accumulate (Fenn and Dunn, 1989).  The accumulation of soil organic matter from increased
13      leaf litter, even in the absence of acidic deposition, can lower soil pH (Binkley, 1992). Lower
14      soil pH can promote loss of nutrients, particularly  cations, from the system, further reducing
15      nutrient availability to the plant. Complex organic compounds in decomposing litter may also
16      tie up nutrients, rendering them less available to plants.
17           Other factors such as nutrient deposition also affect the degree to which O3 effects C and
18      nutrient flow through ecosystems.  In high-pollution sites, the effect of N deposition is difficult
19      to separate from O3 exposure, and reviews of the effect of acidic (and N) deposition on
20      ecosystem nutrient dynamics are important to consider (see Binkley, 1992; Fenn et al., 1998;
21      2003).  For example, at moderately high pollution sites, foliar content  of N is higher than that at
22      lower pollution site,  but so are P, Mg, and Fe contents (Poth and Fenn, 1998). Although
23      significant changes in foliar tissue  chemistry have occurred in response to long-term pollutant
24      deposition in the Carpathians (Mankovska et al., 2002; Fenn et al., 2002), much of this response
25      is correlated to heavy metal and N  and S deposition. At this point, the contribution of O3
26      exposure alone cannot be isolated without careful between-site comparisons of plant response
27      and understanding the deposit!onal velocities of constituent atmospheric species. The significant
28      effects of plant response along known pollution gradients  are important to consider because
29      heavy metal and N and S deposition has  significantly declined over the last decade, while O3
30      exposures remain high and declines in forest health have been sustained.
        January 2005                             9-300        DRAFT-DO NOT QUOTE OR CITE

-------
 1           Models provide a means to track material flows through ecosystems. Two biogeochemical
 2      models were parameterized to capture long-term effects of O3 exposure, N deposition, and
 3      climate on aponderosapine-dominated site in the eastern San Bernardino Mountains.
 4      Simulated O3 exposure resulted in faster production and turnover of foliage and a shift in carbon
 5      from the canopy (15% reduction) to the forest floor (increase of 50 to 60%) (Arbaugh et al.,
 6      1999).  When O3 exposure was combined with that of N deposition, litter mass exponentially
 7      increased.
 8           The direct effect of O3 exposure on below-ground nutrient dynamics and ecosystem
 9      material flow is poorly understood. Additional research will be necessary to understand spatial
10      and temporal dynamics of nutrient and C flow in ecosystems and to separate the effects of O3
11      from those attributable to N deposition.
12
13      9.7.7  Hydrological and Geomorphological
14           At present, there are no publications on the effects of O3 exposure that are carried through
15      at the ecosystem level to changes in mass water flow, channel morphology, riparian habitat
16      complexity, or sediment movement. It is possible that processes occurring at smaller scales are
17      affecting geomorphological processes in ecosystems; however,  difficulty in scaling these
18      responses spatially and temporally have made it difficult to show experimentally. It is possible
19      that O3 exposure affects water quality through changes in energy and material flows, as
20      discussed previously.
21
22      9.7.8  Natural Disturbance Regimes
23           There has been little research on how natural disturbances interact with O3 to affect
24      performance of plants, communities, and ecosystems.  The frequency, intensity, extent, and
25      duration of natural disturbances are variable and unpredictable.  However, there have been
26      enough ecophysiological studies to suggest that O3 could predispose plant communities to
27      certain natural stresses, e.g., drought, stress, or extreme low temperature stress during the winter.
28           While several studies have shown that drought stress reduces O3 uptake through stomatal
29      closure, evidence also suggests that O3 can alter plant water use and susceptibility to drought.
30      In controlled studies, Reich and Lassoie (1984) showed that relatively low O3 concentrations
31      could diminish stomatal control and alter water use efficiency.  Ash trees (Fraxinus excelsior}

        January 2005                              9-301       DRAFT-DO NOT QUOTE OR CITE

-------
 1      exposed to elevated O3 had greater water use early in the growing season, but less water use late
 2      in the growing season when exposed to elevated O3 (Wiltshire et al., 1994).  Under moderate
 3      drought stress, Picea abies trees grown under elevated O3 consumed water faster and showed
 4      higher stomatal conductances than controls (Karlsson et al., 1995).  Pearson and Mansfield
 5      (1993) showed that successive O3 episodes disrupted stomatal function, making Fagus sylvatica
 6      seedlings more susceptible to drought.  Previous year O3 exposure was shown to have a
 7      carryover effect in the following growing season for Fagus sylvatica (Pearson and Mansfield,
 8      1994).
 9           Few studies showing the effects of O3 on water relations of field-grown trees are found in
10      the literature.  However, Grulke et al. (2003a) examined the effects of O3 on canopy transpiration
11      of Pinus Jeffreyi from mesic and xeric microsites and found that trees from mesic sites had 20%
12      more O3 uptake than those in the xeric sites.  The authors also concluded that the mesic trees had
13      greater O3 injury as evidenced by lower needle retention, whereas trees in xeric microsites had
14      greater chlorotic mottle. Chlorotic mottle induced by stomatal uptake of O3 is indistinguishable
15      from that of endogenously produced oxidants resulting from partially closed stomata, a reduction
16      of CO2 inside  the leaf, and production of strong oxidizers within the chloroplast when excited
17      electrons are passed to O2 instead of CO2 under high light levels.
18           Trees living near the limits of their freezing-tolerance range may be especially susceptible
19      to predisposition of freezing injury by O3 (Sheppard et al., 1989). However, Pinus halepensis
20      exposed to elevated O3 had enhanced winter hardiness (Wellburn and Wellburn, 1994). As with
21      the seasonal carryover of drought susceptibility, the influence of elevated O3 on freezing
22      tolerance is carried over from summer to winter. Such effects have been demonstrated for Picea
23      sitchensis (Lucas et al., 1988) and for Picea rubens (Waite et al., 1994).  Sorting out the role of
24      elevated O3 in contributing to frost or low-temperature damage in forests remains  difficult due to
25      the presence of other factors that may affect senescence.
26
27      9.7.9  Scaling to Ecosystem Levels
28           The vast majority of literature describing O3 effects comes from short-duration herbaceous
29      plant or tree seedling studies under controlled conditions. Scaling results from these studies
30      requires extrapolation over both space and time in order to understand the full extent of changes
31      in ecosystems. In addition to spatial, temporal, and age-related complexities, ecosystems are

        January 2005                              9-302       DRAFT-DO NOT QUOTE OR CITE

-------
 1      composed of organisms whose lifetimes range from hours to centuries (Laurence and Andersen,
 2      2003). Forested ecosystems are affected by environmental conditions such as water and nutrient
 3      availability, as well as by intra- and interspecies competition.  Therefore, direct experimentation
 4      to determine the response of forested ecosystems is not simply a matter of determining the effect
 5      of O3 on individual mature trees.  In  addition, even if an experiment can be conducted,
 6      extrapolation of the results across landscapes and regions remains challenging. Nonetheless,
 7      models provide a means to explore possible long-term changes and to identify important
 8      research uncertainties.
 9           Approaches to scaling fall roughly into two categories: (1) process-based modeling to
10      extrapolate physiological responses to O3 based on seedling studies and (2) field assessments
11      using surveys and growth correlations, often in association with  stand-level models to address
12      ecosystem complexity. Comparatively good information is available on process level effects  of
13      O3 in seedlings and, therefore, some  models offer the opportunity to use this information to scale
14      O3 effects at the stand and regional scales (Hogsett et al., 1997; Fuhrer et al., 1997; Chappelka
15      and  Samuelson, 1998; Laurence et al., 2000).
16
17      9.7.9.1   Scaling from Seedlings to Mature Trees
18           A number of investigators have used simulation models based on physiological processes
19      to integrate available data and predict the effects of O3 on mature trees.  Such models predict tree
20      growth by simulating fundamental mechanisms, rather than through a statistical analysis of
21      empirical data. For instance, the process of photosynthesis is simulated based on environmental
22      conditions and physiological characteristics, and then the fixed C is allocated to plant growth
23      using principles of plant physiology. Models based on mechanisms should be applicable across
24      wide areas if the important functional relationships are represented accurately in the models and
25      if the environmental conditions are accurately identified. The ability of six models (TREGRO,
26      CARBON, ECOPHYS, PGSM, TREE-BGC, and W91) to simulate the effects of climatic
27      change and O3 have been reviewed by Constable and Friend (2000).  Of these models, only
28      PGSM and TREGRO explicitly simulated the effects of O3 on foliar processes.
29           The TREGRO model was used to simulate C allocation and tissue growth in seedlings and
30      mature red oak trees based on the experimental data discussed above (Weinstein et al.,  1998).
31      For  seedlings at 2x-ambient O3, only the total nonstructural carbohydrate (TNC) storage pool

        January 2005                              9-303        DRAFT-DO NOT QUOTE OR CITE

-------
 1      was predicted to be affected. For mature trees, large decreases were predicted for TNC, leaves,
 2      stem, branch, and both fine and coarse roots. Most predicted effects in mature trees were
 3      consistent with observations in the field, but the simulations overestimated the effect of
 4      2x-ambient O3 on root TNC and growth.  The authors suggested that this discrepancy may
 5      have been due to trees reducing respiration in response to O3 stress, a response not simulated in
 6      the model.
 7           For Abies concolor, TREGRO was parameterized and simulated growth of a mature tree
 8      for 3 years to test for effects of O3 exposure and drought  stress (Retzlaff et al., 2000).
 9      Reductions in O3 exposure-mediated carbon assimilation were translated to losses in whole tree
10      biomass that probably would not be detectable in the field.  However, TNC levels in branch
11      tissue were simulated to be lowered by over 50%, and branch growth was reduced in a
12      moderately polluted site relative to  a clean site. Low O3  exposure (sufficient to decrease C
13      assimilation by 2.5%) and drought stress (25% reduction in annual precipitation, which is
14      common on a decadal scale) acted synergistically to reduce C gain of whole tree biomass of
15      A. concolor.  Simulated results of the tests were comparable to effects found in OTCs for
16      seedlings and pole-sized trees in clean and moderately polluted sites.
17           Models such as TREGRO are usually parameterized from many different sources of data,
18      including chamber experiments and plantations, from seedlings to mature trees, making it
19      difficult to validate that they reproduce changes that occur  as trees develop from seedlings to
20      maturity. To address this issue, physiological and growth data were collected from a natural
21      stand of P.  ponderosa and used to parameterize the TREGRO model (Grulke and Retzlaff,
22      2001). Representative trees of each of five tree age classes were selected based on population
23      means of morphological, physiological, and nearest neighbor attributes.  Seedlings were
24      observed to differ significantly from pole-sized and older trees in most physiological traits. The
25      changes in biomass with tree age predicted from the model closely matched those of trees in the
26      natural stand.
27           The PGSM model was used to simulate Pinus ponder osa seedling growth responses to O3
28      exposure and drought stress (Chen et al.,  1994). Drought stress was predicted to reduce the
29      effect of O3 on growth, as was observed in the experimental data.  The TREGRO model was
30      used to simulate responses of Pacific Coast and interior varieties of P. ponder osa to five
31      simulated O3  exposures between subambient and 3 x-ambient (Constable and Taylor, 1997).

        January 2005                             9-304       DRAFT-DO NOT QUOTE OR CITE

-------
 1      Simulated growth of var. ponderosa was reduced more than that in var. scopulorum with all O3
 2      exposures.  Drought was protective of O3 exposure.  Similar results were also found in a
 3      relatively moist ponderosa pine plantation (Panek and Goldstein, 2001), whereas drought was
 4      synergistically deleterious with cumulative O3 exposure in a natural stand (Grulke et al., 2002b).
 5           For Pinus taeda, the Plant Growth Stress Model (PGSM) was calibrated with seedling data
 6      and then used to simulate the growth of mature trees over a 55 year period in the Duke Forest,
 7      NC, using estimates of historical O3 concentrations (Chen et al., 1998).  Simulated stem diameter
 8      and tree height were comparable to observed values. In another simulation using TREGRO,
 9      loblolly pine was more sensitive (greatest reduction in C gain) to a peak O3 episode in July
10      (Constable and Retzlaff  1997), whereas mature tulip poplar (Liriodendron tulipiferd) was more
11      sensitive to a peak O3 episode in August.
12           For Populus tremuloides, the ECOPHYS model was used to simulate the relative  above-
13      ground growth response  of an O3-sensitive clone (259) exposed to square-wave variation in O3
14      concentration (Martin et al., 2001).  The model adequately simulated several effects of O3,
15      including a greater effect on stem diameter than on stem height, earlier leaf abscission,  and
16      reduced stem and leaf dry matter production at the end of the growing season. For Acer
17      sacchamm, the TREGRO model was use to predict effects of a 10-year O3 exposure on root and
18      stem growth of a simulated 160-year-old tree (Retzlaff et al., 1996). Twice-ambient O3 exposure
19      (for Ithaca, NY) was predicted to deplete the TNC pools and reduce fine root production.
20
21      9.7.9.2 Surveys, Growth Correlations and Stand-Level Modeling
22           Stand-level studies have included surveys  of O3 symptoms, correlations of radial growth
23      with O3 and other environmental factors, and regional scale modeling.  In addition, open air O3
24      exposure systems, such as those being used on Populus tremuloides-mixed stands in northern
25      Wisconsin  (Karnosky et al., 2003a) and on Fagus sylvatica and Picea abies in Germany (Nunn
26      et al., 2002) offer an opportunity to examine larger plot sizes, older trees, and trees growing
27      under realistic competition. Plots along natural  O3 gradients, as have been used very effectively
28      in southern California forest studies (Miller and McBride, 1999) and with P. tremuloides stands
29      in the Great Lakes region (Karnosky et al., 1999), offer additional insights into ecosystem level
30      responses.  Undoubtedly, however, simulation modeling will have to become an integral
31      component of research in order to predict adequately ecosystem responses to O3 (Laurence and

        January 2005                             9-305       DRAFT-DO  NOT QUOTE OR CITE

-------
 1      Andersen, 2003). Results of these approaches are discussed below, organized into three U.S.
 2      regions: (1) northern states (including the upper Midwest and the Northeast), (2) southeastern
 3      states, and (3) western states (primarily California). A fourth section contains selected
 4      information from Europe.
 5
 6           Northern and Midwestern United States. In recent years, the USD A Forest Service has
 7      conducted systematic O3 biomonitoring surveys in most north-central and northeastern states
 8      (Smith et al., 2003; Coulston et al., 2003). Plots are located on a systematic grid, and trained
 9      field crews evaluate up to 30 plants of up to six species that have foliar injury symptoms
10      diagnostic of O3 damage.  For the United States as a whole, injury has been found more often  in
11      eastern than in interior or west-coast states. As expected, O3 injury is more common and more
12      severe in areas with higher O3 concentrations. Of sampled Prunus serotina plots, -12% were
13      estimated to be at high risk for injury based on a injury index derived from the survey data
14      (Coulston et al., 2003).  P. serotina was estimated to be at risk for injury on the Allegheny
15      Plateau and the Allegheny Mountains (in Pennsylvania, West Virginia, and Maryland), as well as
16      in the coastal plain  of Maryland and Virginia.
17           Ozone concentration, foliar injury, and physiological traits were measured on P. serotina
18      trees of different sizes in Pennsylvania (Fredericksen et al., 1995). The proportion of foliage
19      injured was 46% for seedlings,  15% for saplings, and 20% for canopy trees. Cumulative O3 flux
20      was the most useful O3 metric for predicting injury. Injury  was negatively correlated (r2 = 0.82)
21      with net photosynthetic  rates, but was not related to stomatal conductance. P. serotina is
22      discussed further below, because detailed surveys have been conducted in the Shenandoah and
23      Great Smoky Mountains National Parks in the southeastern United States.
24           Over the past several decades, some surveys ofPinus strobus have reported significant
25      associations between foliar injury and reduced growth (Anderson et al., 1988; Benoit et al.,
26      1982).  However, a review of 93 surveys conducted from 1900 through the late 1980s concluded
27      that methodological problems were pervasive, including such issues as proximity to roads, lack
28      of peer review, lack of random sampling, small sample sizes,  and lack of quantitative methods to
29      estimate severity (Bennett et al., 1994). Because  of these problems, along with evidence of
30      adequate growth rates for P.  strobus regionally and contradictory evidence from numerous
31      studies of symptom production in response to controlled O3 exposure, these authors concluded

        January 2005                              9-306        DRAFT-DO NOT QUOTE OR CITE

-------
 1      that there was no clear evidence of decline in P. strobus (Bennett et al., 1994). A more recent
 2      study in Acadia National Park (ANP) in Maine found no association between O3 exposure in
 3      OTCs and symptom development in P. strobus, calling into question whether symptoms
 4      previously ascribed to O3 may be caused by some other stress (Kohut et al., 2000). However,
 5      another ANP study found significant correlations between O3 exposure and the radial growth of
 6      trees during 10 years in 7 of 8 stands examined (102 trees total; Bartholomay et al., 1997).
 7      Taken together, these results suggest that there may not an association between growth of P.
 8      strobus trees and putative O3 symptoms, but there may be an association between O3 exposure
 9      and radial growth of mature trees in the field.
10           The study of O3 effects was undertaken from 1990 to 1993 in the ANP in Maine, because
11      this location experiences elevated O3 exposures due to transport from urban areas located upwind
12      (Kohut  et al., 2000).  Thirty-two species of plants found in the park were propagated and
13      exposed to O3 in OTCs.  In addition, ambient O3 concentrations were monitored at the study site
14      at 15 m above sea level and near the top of Cadillac Mountain at 470  m above sea level. At the
15      study site, the maximum 1-h O3 concentration was 140 ppb, which occurred in both 1990 and
16      1991. Daytime 12-h O3  concentrations were 35, 41,  36, and 37 ppb during the four years; and O3
17      concentrations were consistently higher at the high-elevation site.  Species showing foliar injury
18      at ambient O3 concentrations included Prunus serotina, Populus tremuloides, Fraxinus
19      americana, Pinus banksiana, big-leaf aster, and spreading dogbane. Species showing foliar
20      injury at 1.5 x ambient O3  concentrations included Betula populifolia, small sundrops, and
21      bunchberry. Species remaining uninjured at 2 x ambient O3 concentrations included Betula
22     papyri/era., Pinus strobus,  Pinus rigida, Picea rubens, Thuja occidentalis, Quercus robur,
23      Canada bluejoint grass, wild radish, and Canada mayflower. Because of their O3  sensitivity and
24      diagnostic symptoms, big-leaf aster, spreading dogbane, Populus tremuloides, Fraxinus
25      americana, and Prunus serotina were recommended as bioindicators  for the ANP.
26           The PnET-II model was applied to 64 locations across the northeastern United States to
27      simulate the effects of ambient O3 on mature hardwood forests (Ollinger et al., 1997).  In this
28      model, O3 effects on each of several layers of the forest canopy were represented by a single
29      linear equation relating predicted O3 uptake to decreased net photosynthetic rate.  Wood growth
30      was predicted to decrease between 3 to 22%, with greatest reductions in southern portions of the
31      region where O3 levels were highest and on soils with high water-holding capacity where

        January 2005                             9-307       DRAFT-DO NOT QUOTE OR CITE

-------
 1      drought stress was absent. Little variation was predicted among years, because high O3 often
 2      coincided with hot, dry weather conditions that reduced predicted stomatal conductance and
 3      O3 uptake.
 4           In order to estimate the impact of O3 on forests, effects must be evaluated not only on
 5      individuals, but also on mixtures of species and the composition of forest stands.  The PnET
 6      model described above evaluated the effects of O3 on broad forest types (an evergreen/deciduous
 7      mix), but did not address specific forest species composition. In order to address competition
 8      among species, the TREGRO model was linked to the ZELIG forest stand growth model and a
 9      geographic information system was created to predict the effects of O3 across the north-central
10      and northeastern United States (Laurence et al., 2000). ZELIG is a gap-succession model used
11      to simulate succession in mixed stands typical of eastern and northern forests.  Ambient O3
12      generally caused a reduction of 2 to 4% in the growth ofQuercus robur across the region during
13      the 100-year simulation.  The response followed the pattern of O3 exposure, with little effect in
14      the northwest part of the region, but with greater  effect in southern locations.  The O3 response of
15      Acer saccharum to O3 varied widely, but the overall growth response was always positive,
16      indicating that the evergreen/deciduous  mix was able to take advantage of the  decrease in the
17      growth of Q. robur and other species caused by O3.  In the northernmost part of the region,
18      A. saccharum growth increased by up to 3%, but in the southern part of the region, its growth
19      increased by up to 12%.  The authors ascribed this enhanced growth to a combination of warmer
20      temperatures and reduction in the growth of Primus serotina, a minor component of the
21      simulated stand that was very sensitive to O3.
22
23           Southeast  United States. In a survey of the Great Smoky Mountains National Park, foliar
24      injury attributed to O3 was found on 47% of the more than 1,600 plants examined (Chappelka
25      et al., 1997). In  subsequent surveys of injury in the park, injury was found on  mature trees of the
26      following species: Sassafras albidum, Prunus serotina, and Liriodendron tulip/era (Chappelka
27      et al., 1999a,b).  In a similar study in Shenandoah National Park, injury was found on Fraxinus
28      americana, P. serotina, andZ. tulip/era (Hildebrand et al.,  1996).
29           For Prunus serotina seedlings grown in soil in OTCs and exposed to relatively low
30      ambient levels of O3 in Pennsylvania, there was no correspondence between visible foliar stipple,
31      leaf gas exchange, and seedling growth between two families previously shown to differ in O3

        January 2005                             9-308        DRAFT-DO NOT QUOTE OR CITE

-------
 1      symptoms (Kouterick et al., 2000). However, significant exposure-response relationships were
 2      found for foliar injury in the Great Smoky Mountains and Shenandoah National Parks.  In each
 3      park, foliar injury was evaluated on mature P. serotma trees on three plots at different elevations
 4      near O3 monitors during 1991 to 1993 (Hildebrand et al., 1996; Chappelka et al., 1999a,b).
 5      In 1991, incidence was 60% and 45% for the two parks and 33% in both parks during 1992 and
 6      1993.  Symptoms were greater at the highest elevations where O3 concentrations were highest.
 7      In another study, radial growth rates were measured in 44 P. serotma trees ranging in age from
 8      19 to 56 years old with and without O3 symptoms, at three sites in the Great Smoky Mountains
 9      National Park. Trees with O3 symptoms were compared to similar-sized trees with few
10      symptoms. There was no evidence that trees with O3 symptoms had lower growth rates (p = 0.6)
11      (Somers et al., 1998).
12          In the Great Smoky Mountains National Park, radial growth rates were measured for
13      44 L. tulip/era trees ranging in age from 30 to 58 years old, with and without O3  symptoms, at
14      three sites at different elevations (Somers et al., 1998).  Trees with O3 symptoms averaged 30%
15      lower growth rates  over ten years (p = 0.0005).  Seedlings of Liriodendron tulip/era were
16      exposed for two seasons to 2x-ambient  O3 exposures in OTCs in Delaware, OH (seasonal
17      SUMOO exposures of 107 and 197 ppmh) (Rebbeck, 1996).  Foliar O3 symptoms were observed,
18      but growth was not reduced.
19          In order to evaluate the influence of interspecies competition on O3 effects, the linked
20      TREGRO and ZELIG modeling system was used to predict the effects of O3 over 100 years on
21      the basal area of species in a Liriodendron tulipfera-dominated forest in the Great Smoky
22      Mountains National Park (Weinstein et  al., 2001).  Ambient O3 was predicted to  reduce the basal
23      area of L. tulip/era by 10%, whereas a 1.5x-ambient exposure was predicted to cause a 30%
24      reduction.  Basal area of Acer rubrum and Prunus serotma was predicted to increase for some
25      years, but then decrease by up to 30%, with few changes in the total basal area of all species by
26      the end of the simulation.
27          In order to evaluate the influence of interspecies competition on O3 effects, the linked
28      TREGRO and ZELIG modeling system was used to predict the effects of O3 on the basal area of
29      Pinus taeda and Liriodendron tulip/era  growth throughout their ranges (Laurence et al., 2003).
30      The models were parameterized using biological and meteorological data from three sites in the
31      southeastern United States (in Alabama, Louisiana, and North Carolina). Forest  stand response

        January 2005                             9-309       DRAFT-DO NOT QUOTE OR CITE

-------
 1      to five O3-exposure regimes with annual SUM06 values ranging from 0 to 100 ppmh per year
 2      was simulated for 100 years. The simulated basal area of the two species was generated within
 3      the context of four other tree species common in southeastern forests. Basal area of P. taeda was
 4      highly responsive to precipitation and O3 exposure, with the greatest increases under high-
 5      precipitation, lowO3-exposure scenarios and the greatest decreases under low-precipitation, high
 6      O3-exposure scenarios. The basal area of L. tulipifera did not significantly differ (+10%) from
 7      simulations using a "base case" (ambient O3, average precipitation).
 8           Systematic biomonitoring surveys found that approximately 24% of sampled sweetgum
 9      and 15% of sampled Pinus taeda plots were estimated to be at high risk for foliar injury on the
10      coastal plain of Maryland and Virginia (Smith et al., 2003;  Coulston et al., 2003). In a study in
11      Tennessee, the effect of ambient (uncontrolled) O3 on 28 mature canopy-dominant 50 to
12      90-year-old P. taeda trees in five stands was measured over a 5-year period (McLaughlin and
13      Downing, 1995, 1996). Of many O3 metrics, a 3-day average of hourly O3 values > 40 ppb
14      (AOT40) was found to best explain short-term variation in  stem expansion as measured with
15      dendrometer bands. Interactions  between O3, temperature,  and  drought stress (as indicated by
16      the weekly moisture stress index) accounted for 63% of the short-term variation in stem growth
17      rates. Because there are interactions among O3,  drought stress,  and temperature that may  differ
18      with the averaging time (days to years), this type of study cannot provide conclusive  proof of
19      cause and effect (Reams et al., 1995). However, the results do suggest that the effects of O3
20      measured on loblolly seedlings may also be occurring in mature trees in both wet and dry sites.
21      The magnitude of effects of O3 on growth, including interactions with other variables in this
22      study, ranged from 0 to 15% over 5 years, with an average  of 5.5%.
23
24           Western United States.  The USD A Forest Service conducted O3 biomonitoring surveys in
25      Washington, Oregon, and California during one year (1998), and in the  Sierra Nevada and
26      Sequoia National Forest every other year for several decades (Campbell  et al., 2000). Overall,
27      only one plot showed any symptoms of O3 injury outside of the Sierra Nevada and Sequoia
28      National Forests.  In the Sierra Nevada National Forest, between 30 and 40% of trees showed
29      injury from 1989 through 1997.  In  the Sequoia National Forest, between 40 and 50% of the
30      trees surveyed showed injury from  1990 through 1998.
        January 2005                             9-310        DRAFT-DO NOT QUOTE OR CITE

-------
 1           For Pinus ponderosa along a well-studied gradient of O3 exposure in the San Bernardino
 2      mountains, chlorotic mottle was highest on foliage at the most polluted site, as has been found
 3      previously (Grulke and Balduman, 1999). Based on whole-tree harvests, root biomass was
 4      lowest at the most polluted sites, confirming previous studies with seedlings under controlled
 5      conditions, as discussed above.  Ozone responses in highly polluted environments such as
 6      Southern California may not be predicted adequately by extrapolating effects from single-factor
 7      experiments. Instead, combined approaches utilizing field experiments and modeling efforts
 8      may be required to properly account for a combination of stressors including O3, N deposition,
 9      and drought. Furthermore, the available studies underscore the lack of correlation between O3
10      symptoms and mature tree effects. If a field survey fails to find a correlation between mature
11      tree growth and O3, this result may be due to the dominant effect of another factor such as N
12      deposition and may not be evidence that O3 does not reduce the growth of mature trees.
13           The TREGRO model was used to evaluate how projected future temperature and CO2
14      concentrations might affect the response of individual ponderosa pine to O3 at seven sites in
15      California, Oregon, and Washington (Tingey et al., 2001). As expected, growth decreased with
16      increasing O3 exposure.  Differences in O3 response among sites appeared to be due primarily to
17      differences in precipitation.
18           Often air quality standards do not translate directly into measureable improvements in tree
19      growth or productivity.  To evaluate whether past improvements in air quality have improved
20      ponderosa pine growth,  TREGRO was used to simulate growth at sites in the San Bernardino
21      Mountains in California (Tingey et al., 2004).  Ozone and meteorogical data from the past
22      37 years was used to run the simulations. Despite variation in precipitation and temperature,
23      O3 was found to  reduce  simulated tree growth. The authors were able to simulate growth
24      improvements as air quality improved during the 1980s and 1990s, suggesting that
25      improvements in emission control strategies benefited ponderosa pine. The model simulations
26      were qualitatively consistent with improvements in canopy condition that were observed at sites
27      where O3 reductions were the greatest.
28
29           Studies in Europe. In a 4-year study ofFagus sylvatica in Switzerland at 57 forest sites
30      ranging in age from 65 to 173 years, stem increment was found to decrease with increasing
31      maximum O3 exposure (Braun et al., 1999).  In this study, O3 concentration was estimated by

        January 2005                             9-311        DRAFT-DO NOT QUOTE OR CITE

-------
 1      interpolation among monitoring stations, and other site conditions such as soil water status and
 2      temperature were interpolated from weather stations.  Other factors such as N deposition, tree
 3      diameter, and canopy dominance were also found to be significantly associated with stem
 4      increment.  The maximum annual O3 dose (expressed as AOT40) was found to be more strongly
 5      associated with decreased stem increment than was the average O3 dose over the 4 years.
 6      A growth reduction of 22.5% (confidence interval 14.3 to 28.6%) was associated with each
 7      10 ppmh increment of O3 (expressed as AOT40). This decrease was steeper than the 6.1%
 8      growth reduction summarized previously from several OTC studies with F. sylvatica seedlings
 9      (Fuhrer et al., 1997).  However, the authors suggest that this difference may be explained largely
10      by the 4 years of exposure in their forest survey study as compared to the one-year exposures for
11      seedlings. As with any forest survey, these results must be interpreted with caution because O3
12      exposure was correlated with other variables, such as tree age and the deposition of NO2
13      and SO2.
14
15      9.7.10  Summary of Ecological Effects of Ozone Exposure on
16              Natural Ecosystems
17           In this chapter, an effort has been made to discuss the adverse effects of O3 on natural
18      ecosystems within the context of the SAB framework for assessing and reporting ecological
19      conditions (Young and Sanzone, 2002). Using this framework, there is evidence that
20      tropospheric O3 is an important stressor of natural ecosystems, with well-documented impacts on
21      the biotic condition, ecological processes, and chemical/physical nature of natural ecosystems
22      (Figure 9-20). In turn, the effects of O3 on individual plants and processes are scaled up through
23      the ecosystem affecting processes such as energy and material flow, intra and interspecies
24      competition, and NPP.  Thus, effects on individual keystone species and their associated
25      microflora and fauna may cascade through the ecosystem to the landscape level.  This suggests
26      that by affecting water balance, cold hardiness, tolerance to wind, and by predisposing plants to
27      insect and disease pests, O3 may even influence the occurrence and impact of natural
28      disturbance.  Despite the probable occurrence of such effects, however, there are essentially no
29      instances where highly integrated ecosystem-level studies have conclusively  shown that O3 is
30      indeed altering ecosystem structure and/or function.
31

        January 2005                            9-312        DRAFT-DO NOT QUOTE OR CITE

-------
                    Hydrolie Alteration
                    Habitat Conversion
                    Habitat Fragmentation
                    Climate Change
                    Invasive Non-native Species
                    Turbidity/Sedimentation
                    Pesticides
                    Disease/Pest Outbreaks
                    Nutrient Pulses
                    Metals
                    Dissolved Oxygen Depletion
                    Ozone (Tropospheric)
                    Nitrogen Oxides
                    Nitrates
   Hydroiic Alteration
   Habitat Conversion
   Habitat Fragmentation
   Climate Change
   Over-Harvesting Vegitation
   Large-Scale Invasive
      Species Introduction
   Large-Scale Disease/Pest
      Outbreaks
                             Hydroiic Alteration
                             Habitat Conversion
                               Climate Change
                      Over-Harvesting Vegitation
                         Disease/Pest Outbreaks
                            Altered Fire Regime
                           Altered Flood Regime
                          Hydroiic Alteration
                          Habitat Conversion
                          Climate Change
                          Turbidity/Sedimentation
                          Pesticides
                          Nutrient Pulses
                          Metals
                          Dissolved Oxygen Depletion
                          Ozone (Tropospheric)
                          Nitrogen Oxides
                          Nitrates
                          Sulfates
                          Salinity
                          Acidic Deoosition
t/J
£
+•>
w
Hydroiic Alteration
Habitat Conversion
Climate Change
Pesticides
Disease/Pest Outbreaks
Nutrient Pulses
Dissolved Oxygen Depletion
Nitrogen Oxides
Nitrates
Sylfates
                           Hydroiic Alteration
                           Habitat Conversion
                        Habitat Fragmentation
                             Climate Change
                       Turbidity/Sedimentation
        Figure 9-20.   Common anthropogenic stressors and the essential ecological attributes
                        they affect.

        Source:  Modified from Young and Sanzone (2002).
1             Systematic injury surveys demonstrate that foliar injury occurs on sensitive species in

2       many regions of the United States.  However, the frequent lack of correspondence between foliar

3       symptoms and growth effects means that other methods must be used to estimate the regional

4       effects of O3 on tree growth rates.  Investigations  of the radial growth of mature trees in

5       combination with data from many controlled studies with seedlings, as well as a few studies with
        January 2005
     9-313
                DRAFT-DO NOT QUOTE OR CITE

-------
 1     mature trees suggest that ambient O3 is reducing the growth of mature trees in some locations.

 2     Studies using models based on tree physiology and forest stand dynamics suggest that modest

 3     effects of O3 on growth may accumulate over time and may interact with other stresses. For

 4     mixed-species stands, such models predict that overall stand growth rate is generally not likely to

 5     be affected. However, competitive interactions among species may change as a result of growth

 6     reductions  of sensitive species. These results suggest that O3 exposure over decades may be

 7     altering the species composition of forests in some regions.
 8

 9     RESEARCH NEEDS

10           The knowledge base for examining the range of ecological effects of O3 on natural

11     ecosystems is growing, but significant uncertainties remain  regarding O3 effects at the ecosystem

12     level.  For  example, there is a need for information on the following ecosystem-level responses:

13       •  Ecosystem Processes.  Little is known about the effects of O3 on water, C, and nutrient
            cycling, particularly at the stand and community levels. Effects on belowground
            ecosystem processes in response to O3 exposure alone and in combination with other
            stressors are critical to projections  at the watershed and landscape scales.  Little is yet
            known about the effects of O3 on structural or functional components of soil food webs,
            or how these impacts could affect plant species diversity (Andersen, 2003).

14       •  Biodiversity and Genetic Diversity. The study of genetic aspects of O3 impacts on natural
            ecosystems has been largely correlational in nature and it remains to be shown
            conclusively whether O3 affects biodiversity or genetic diversity (Pitelka, 1988; Winner
            et al.,  1991; Davison and Barnes, 1998). Studies of competitive interactions under
            elevated O3 levels are needed (Laurence and Andersen, 2003), and reexamination via new
            sampling of population studies to bring in a time component to previous studies showing
            spatial variability in population responses to O3 are needed.  These studies could be
            strengthened by modern molecular methods to quantify impacts on diversity.

15       •  Natural Ecosystem Interactions with the Atmosphere. Little is known  about feedbacks
            between O3 and climate change on  volatile organic compound (VOC) production, which in
            turn, could affect O3 production (Fuentes et al., 2001).  At moderate-to-high O3 exposure
            sites, aberrations in stomatal behavior could significantly affect individual tree water
            balance of sensitive trees, and if the sensitive tree species is dominant, hydrologic balance
            at the  watershed and landscape level could be affected. This has not been addressed in any
            model because O3 exposure effects, if included in the modeling effort have assumed a
            linear relationship between assimilation and stomatal conductance.

16       •  Below-Ground Inter actions. While the negative effects of O3 on below ground growth are
            well characterized, interactions of roots with the soil or microorganisms are not.
       January 2005                              9-314       DRAFT-DO NOT QUOTE OR CITE

-------
            Other Interactive Effects.  Interaction studies with other components of global change
            (e.g.,warming, increasing atmospheric CO2, N deposition, etc.) or with various biotic
            stressors are needed to better predict complex interactions likely in the future (Laurence
            and Andersen, 2003). Whether O3 will negate the positive effects of an elevated CO2
            environment on plant carbon and water balances is not yet known; nor is it known if these
            effects will scale up through the ecosystem.  How might O3 affect the progress of pest
            epidemics and insect outbreaks as concentrations increase is unclear (Skarby et al., 1998).

            Reproduction Effects. Information concerning the impact of O3 on reproductive processes
            and reproductive development under realistic field or forest conditions are needed as well
            as examination of reproductive effects under interacting pollutants (Black et al., 2000).

            Comparative Extrapolation. The vast majority of O3 studies of trees have been conducted
            with young, immature trees and in trees that have not yet formed a closed canopy.
            Questions remain as to the comparability of O3 effects on juvenile and mature trees
            and on trees grown in the open versus those in a closed forest canopy in a competitive
            environment (Chappelka and Samuelson, 1998; Kolb and Matyssek, 2001; Samuelson
            and Kelly, 2001).

            Scaling-Up Issues.  Scaling the effects of O3 from the responses of single or a few plants
            to effects on communities and  ecosystems is a complicated matter that will require a
            combination of manipulative experiments with model ecosystems,  community and
            ecosystem studies along natural O3 gradients, and extensive modeling efforts to project
            landscape-level, regional, national and international impacts of O3.  Linking these various
            studies via impacts on common research quantification across various scales using
            measures of such factors as leaf area index or spectral reflective data, which could
            eventually be remotely sensed  (Kraft  et al., 1996; Panek et al., 2003), would provide
            powerful new tools for ecologists.

            Comparative Risk Assessment Methodologies. Methodologies to determine the important
            values of services and benefits derived from natural ecosystems such that these could be
            used in comprehensive risk assessment for O3 effects on natural ecosystems (Heck et al.,
            1998).
 8     9.8  ECONOMIC EVALUATION OF OZONE EFFECTS ON
 9          AGRICULTURE, FORESTRY AND NATURAL ECOSYSTEMS

10     9.8.1  Introduction

11          The adverse consequences of ambient of air pollutant exposures on vegetation, ecosystems
12     and components of the material environment have been documented since the beginning of the
13     industrial revolution.  Attempts to quantify the monetary damage and injury resulting from
       January 2005                            9-315       DRAFT-DO NOT QUOTE OR CITE

-------
 1      tropospheric O3 exposures to managed agriculture, forests, and natural ecosystems date back at
 2      least to the 1950s.
 3           Both methodological and data problems plagued early efforts to assess the monetary
 4      damages of air pollution to crops and natural vegetation. Adams and Crocker (1989) discussed
 5      the methodological issues, e.g., a lack of reliable data on effects from air pollutants on crop
 6      yields or the failure to develop and apply appropriate economic models. Some of these problems
 7      were remedied by the EPA's National Crop Loss Assessment Network (NCLAN) in the 1980s.
 8      The EPAs NCLAN facilitated the performance of economic assessments by providing O3-crop
 9      yield data with which to estimate O3-crop yield response functions (see Heagle [1988] for a
10      review of NCLAN procedures and findings).  NCLAN also funded a series of economic
11      assessments that, along with subsequent economic assessments, documented substantial
12      economic damages to agriculture. (See Spash [1997] for a detailed review of economic
13      assessments, many of which used NCLAN data.)
14           Since the completion of the NCLAN program in the late 1980s, the number of economic
15      assessments of air pollution studies focusing on terrestrial ecosystems in general, and agriculture
16      in particular, has declined. For example, for the period of 1980 to 1990, 33 economic studies of
17      O3 and other air  pollutant effects on U.S. crops were published in peer-reviewed journal outlets
18      (Spash, 1997). However, in preparing this section of the current criteria document, only four
19      peer-reviewed economic assessments were found for the decade of 1991 to 2000 that addressed
20      vegetation in the United States.  In addition, one peer-reviewed article (Kuik et al., 2000) was
21      found dealing with agriculture in the Netherlands. Recent interest in global climate change and
22      the potential effects of global warming on O3 and other photochemical oxidants, has renewed
23      interest in the effects of air pollution on both managed and unmanaged terrestrial ecosystems
24      (Adams et al.,  1998). In addition, concern is growing for regarding the effects of air pollutants
25      on natural ecosystems and on the services they provide (Daily, 1997). Unfortunately, this
26      interest has  not yet translated into additional peer-reviewed publications addressing O3 or other
27      air pollutants effects on ecosystems.
28           This section of the current criteria document first discusses the availability of economic
29      information and  its usefulness in forming environmental policy.  Next,  economic assessments of
30      air pollution effects and findings from the 1996 AQCD (U.S. Environmental Protection Agency,
31      1996) are discussed, followed by a synthesis of the limited literature available since the 1996

        January 2005                              9-316        DRAFT-DO NOT QUOTE OR CITE

-------
 1      AQCD with respect to O3 effects on agriculture, forestry, and ecosystems. Finally, limitations
 2      and continuing uncertainties are reviewed. The most fundamental of these is the lack of
 3      measurements of the economic effects of air pollution on natural ecosystems. Other issues
 4      include the variability of performance in both managed and natural ecosystems under increased
 5      climatic and air pollution variability as well as the challenges related to spatial and temporal
 6      scales used in performing economic assessments.  To date, this set of effects has been sparsely
 7      addressed.
 8
 9      9.8.2 The Measurement of Economic Information
10           Economic science is an exercise in deductive logic in which testable hypotheses about the
11      behavior of economic agents (i.e., farmers, consumers, resource owners) and markets are
12      deduced from a body of theory.  That body of theory is based on a series of premises proposed
13      by economists and philosophers dating back over two centuries to Adam Smith.  These premises
14      gradually evolved into a theoretical foundation primarily dealing with microeconomics,
15      culminating in structural relationships that define the operation of markets. This foundation was
16      first laid out in a comprehensive and rigorous fashion by Alfred Marshall in 1920.  Paul
17      Samuelson (1948) formalized these theoretical relationships, resulting in what is sometimes
18      referred to as modern, or neoclassical, economics.
19           The insights gained from the theoretical foundation of economics helped shape the nature
20      of applied economic, or policy-relevant research. An example of such applied research is when
21      economists seek to measure the economic consequences of air pollution on agriculture. Such an
22      application is described in Adams and Horst (2003), who provided a graphical representation of
23      the measurement of the effects of air pollution on the well-being of producers and consumers.
24      Economic theory is applied to real-world problems when the methods of economics and the need
25      for data from other disciplines come into play. When measuring the economic effects of an
26      environmental change,  economists need an economic model or method that is theoretically
27      consistent, i.e, defensible, as well as data with which to estimate both economic and
28      environmental science relationships for use in the model. Among accepted economic assessment
29      methods, the actual choice is frequently determined by the nature of the problem to be addressed.
30      It should be noted that the choice of assessment method can affect the type of economic
31      information that is obtained.  Even with a given assessment method, results are sensitive to

        January 2005                             9-317       DRAFT-DO NOT QUOTE OR CITE

-------
 1      specific data treatment or assumptions (Adams, 1999). For example, some methods only
 2      measure effects on a particular group e.g., farmers.  Other methods may measure effects across
 3      several groups. Thus, one should not expect the magnitudes of damage or benefits to be
 4      identical across economic assessments. One should, however, expect that the direction of the
 5      effects will be similar.
 6           Once it is established that an assessment meets basic economic criteria, e.g., including
 7      human behavioral responses, the selection of the specific economic assessment method is often a
 8      relatively minor issue in terms of estimating benefits of air pollution control (or disadvantages of
 9      increases in air pollution). Although results differ across approaches, the differences  are largely
10      attributable to specific features of the assessment (e.g., whether the natural science data include a
11      particular effect or relationship, whether effects on consumers are measured, and so forth).  The
12      nature and quality of the air quality forecasts used in the assessments can greatly influence the
13      sensitivity of the assessments (Adams and Crocker, 1989).  This is particularly noticeable when
14      dealing with forecasts of seasonal air pollution changes (Adams et al., 1988). From the
15      standpoint of providing policy guidance, the differences in economic estimates attributable to the
16      assessment methods are often swamped by uncertainty in the natural and physical science
17      forecasts.  This has also been noted in recent economic assessments of climate change (Adams
18      et al., 1998). In many settings, the quality of economic assessments of air pollution is likely to
19      be improved more by refining the physical and natural science data used in the assessments than
20      by intensive efforts to fine-tune the assessment techniques (Adams, 1999).
21
22      9.8.3  Understanding of Air Pollutants Effects on the Economic Valuation of
23             Agriculture and Other Vegetation in the 1996 Criteria Document
24           Evidence from the plant science literature cited in the 1996 O3 AQCD (U.S. Environmental
25      Protection Agency, 1996) is unambiguous with respect to the adverse effects of tropospheric O3
26      on some types of vegetation. For example, the 1996 AQCD noted that findings from  the EPA
27      multiyear NCLAN program in the 1980s provided rigorous corroboration of at least two decades
28      of previous research and a century of anecdotal observations that showed that O3 at ambient
29      levels caused physical damage to plants in general and to important crop species in particular.
30      Specifically, NCLAN established that ambient O3 levels resulted in statistically significant
31      reductions in yields for some crop species (Heagle et al., 1988).  The 1996 AQCD also assessed

        January 2005                             9-318        DRAFT-DO NOT QUOTE OR CITE

-------
 1      the results of studies regarding O3 effects on crops, forests, and natural vegetation in more detail.
 2      More recent reviews, such as the comprehensive survey of the economic literature on
 3      agricultural effects by Spash (1997) corroborate the synthesis of results reported in the 1996
 4      AQCD.
 5           The number and quality of assessments of the economic consequences of O3 exposures on
 6      vegetation reported in the 1996 AQCD are primarily a function of the state of evidence obtained
 7      from scientific studies in each vegetation category. For example, the plant science evidence
 8      reviewed in the 1996 AQCD concerning effects of O3 exposures on agricultural crops was
 9      reported to be more valid than for individual tree species or plant communities (ecosystems).
10      As a result, most economic assessments discussed in the 1996 AQCD focused on the data
11      obtained from studies of agricultural crops. The economic literature dealing with O3 effects on
12      forest productivity in the 1996 AQCD is sparse.  The few economic assessments of tree or forest
13      effects reported in the 1996 AQCD were confined to evaluations of assumed or hypothetical
14      changes in output, such as board feet of lumber (e.g., Haynes and Adams [1992]).  As noted in
15      the 1996 AQCD,  O3 effects on ecosystems and their services had not been measured in any
16      systematic fashion and no peer-reviewed economic assessments were yet reported.
17           This section first briefly reviews economic assessments drawn from the review in the 1996
18      criteria document. This review is the benchmark against which recent articles are then discussed
19      in the subsequent section.  As was the case in 1996, the discussion of economic valuation of
20      ecosystem effects is generally limited to conceptual and methodological issues, given the
21      continued lack of empirical analyses in this category.
22
23      9.8.3.1   Agriculture
24           In view of the importance of U.S.  agriculture for both domestic and world consumption of
25      food and fiber, reductions in U.S. crop yields could adversely affect human welfare.  The
26      plausibility of this premise has resulted  in numerous attempts to assess, in monetary terms, the
27      losses from ambient O3 exposures or the benefits of O3 control, to agriculture. Twenty-three
28      assessments of the economic effects of O3 exposures on agriculture were reviewed in the  1996
29      AQCD, highlighting key issues in the validity of these assessments (U.S. Environmental
30      Protection Agency, 1996). First, the evidence should reflect how crop yields will respond under
31      actual field conditions to O3 exposures.  Second, the air quality data used to frame current or

        January 2005                             9-319       DRAFT-DO NOT QUOTE OR CITE

-------
 1      hypothetical effects of O3 on crops should represent actual exposures sustained by crops at
 2      individual sites or production areas.  Finally, the assessment methodology into which such data
 3      are entered should (1) capture the economic behavior of producers and consumers as they
 4      adjust to changes in crop yields and prices that may accompany changes in O3 air quality;
 5      (2) accurately reflect institutional considerations, such as regulatory programs and income
 6      support policies (e.g., provisions of federal Farm Bill legislation), that may result in market
 7      distortions; and (3) use measures of well-being that are consistent with economic principles.
 8           Assessments of O3 damage to agricultural crops reported in the 1996 AQCD employed
 9      procedures for calculating economic losses that met the conditions described above. More
10      specifically, the assessments provided 23 quantitative estimates of the economic consequences
11      of exposures to O3 and other air pollutants to agriculture that reflect producer-consumer
12      decision-making processes, associated market adjustments, and some measure of distributional
13      consequences between affected parties. Many of the economic assessments reviewed in
14      previous O3 documents also focused on O3 effects in specific regions, primarily California and
15      the Corn Belt  (e.g., Garcia et al., 1986). This regional emphasis in the earlier literature may be
16      attributed to the relative abundance of data on crop response and  air quality for selected U.S.
17      regions, as well as the importance of some agricultural  regions (such as California) for the U.S.
18      agricultural  economy.
19           Two U.S. national studies described in previous O3  criteria  documents that are worthy of
20      additional comment are Kopp et al. (1985) and Adams et al. (1986). These were judged to be
21      "adequate" in terms of the three critical areas of data inputs in the 1996 AQCD.  Together, it was
22      argued, they provide a reasonably comprehensive estimate of the economic consequences of
23      changes in ambient air O3 levels on agriculture.  Because of their central role in the 1996 criteria
24      document, these two studies are reviewed briefly below.
25           The Kopp  et al. (1985) and Adams et al. (1986) studies indicated that ambient levels of O3
26      were imposing substantial economic costs of ~$3.4 billion (in 2000 U.S. dollars) on agriculture.
27      Both were judged to suffer from several sources of uncertainty, but the document concluded that
28      these possible improvements  in future assessments were not likely to greatly alter the range of
29      agricultural  benefit estimates arising from O3 reductions for several reasons. First, the studies
30      covered about 75 to 80 % of U.S.  agricultural crops (by value). For inclusion of the other 20%
31      to significantly change the estimates would require that their sensitivities to O3 be much greater

        January 2005                              9-320        DRAFT-DO NOT QUOTE OR CITE

-------
 1      than for the crops that have been included to date.  Second, model-sensitivity analyses reported
 2      in past studies indicate that changes in plant exposure response relationships must be substantial
 3      to translate into major changes in economic benefits estimates. For example, it was assumed
 4      unlikely that use of different exposure measures, or inclusion of interaction effects, would
 5      greatly alter the magnitude of the economic estimates.  Third, it was believed that countervailing
 6      effects would mitigate  against large swings in the estimates, e.g., longer O3-exposure periods
 7      may predict greater yield losses, but O3-water stress interactions tend to reduce loss the
 8      yield estimates.
 9           Other national assessments reported in the 1996 AQCD provided general corroboration of
10      the results of Kopp et al. (1985) and Adams et al. (1986). An evaluation of these studies in
11      terms of the adequacy of information from plant exposure studies, and aerometric and economic
12      data was presented in the 1996 AQCD, along with estimates of benefits or damages associated
13      with changes in O3. Most of the studies built on either Kopp et al. (1985) or Adams et al. (1986).
14      A relevant question was whether subsequent studies provided any "surprises" in terms of
15      magnitude of economic effects.
16           Common themes or findings from these  and earlier O3 and other air pollution studies were
17      summarized in two synthesis papers, Adams and Crocker (1989) and Segerson (1991).  The
18      major conclusion is that the agricultural effects of tropospheric O3 at ambient levels impose
19      economic costs to society or, conversely, that  reductions in ambient  O3 should result in societal
20      benefits.
21           Several studies contained in the 1996 AQCD still are  of interest. For example, one finding
22      pertains to the relationship between federal farm programs and air pollution regulations
23      (McGartland, 1987). In each case, the inclusion of farm programs in the economic models
24      resulted in modest reductions in the economic benefits of O3 control due to increased farm
25      program costs. As Segerson (1987) noted, however, it is not clear that these increased costs
26      should be charged against the potential benefits of an O3-regulatory standard, but rather,
27      considered as an additional cost associated with the inefficiencies of the farm program. It should
28      also be noted that the nature of federal farm programs was changed dramatically by Congress in
29      1996 in an attempt to reduce the federal government's role in agriculture. Although more recent
30      federal legislation, such as the 2002 Farm Bill (U.S. Congress, 2002), appears to be restoring the
31      federal government's role in the farming sector, this issue currently is not as important as

        January 2005                              9-321        DRAFT-DO NOT QUOTE OR CITE

-------
 1      suggested by earlier studies, due to the declining reliance on deficiency payments to farmers,
 2      which tend to distort resource allocation.
 3           Another national study (Adams et al., 1988) analyzed economic benefits under a regulatory
 4      alternative involving a seasonal (crop growing season) O3-exposure index measured as a 12-h
 5      mean, instead of hourly levels or percent changes from ambient as reported in earlier studies.
 6      Specifically, a seasonal average of 50 ppb O3 (measured as a 12-h seasonal mean) with a
 7      95% compliance level, is reported in Adams et al. (1988). The result (a $2.9 billion benefit in
 8      2000 dollars) is  similar to the assumed 25% reduction across all regions reported by Adams et al.
 9      (1986). At least one study also combined environmental stressors (e.g., O3, UV-B radiation) in
10      performing economic assessments. Adams and Rowe (1990), using the same model as Adams
11      et al. (1986), reported that a 15% depletion of stratospheric O3 (resulting in a 13% increase in
12      tropospheric O3) would cause an economic loss of ~$1.4 billion in 2000 dollars attributed to the
13      tropospheric O3  increase. Reducing VOCs/NOx motor vehicle emissions by 10% would result in
14      a benefit of ~$0.3 billion, while a complete elimination of motor vehicle emissions would yield a
15      benefit of ~$3.4 billion (1990 dollars). The range of these numbers is consistent with values
16      reported in Adams et al. (1986), Kopp et al. (1985), and  other national-level analyses, i.e.,
17      estimates of from $1.0 to 2.0 billion for reductions in ambient O3 of 25 to 40%.
18
19      9.8.3.2 Forests (Tree Species) and Natural Ecosystems
20           The long-term nature of air pollution effects on perennial species such as trees creates
21      challenges to plant scientists in attempts to sort out the effects of specific individual stressors
22      such as O3 from among the many other potential causal factors (Skelly, 1988). It also creates
23      problems in terms of measuring the impacts on direct economic value, such as reductions in
24      board-feet of lumber produced per unit of time.
25           Most of the literature in the 1996 AQCD dealing with forest species reported the effects of
26      O3 exposures in  terms of foliar injury (Davis and Skelly,  1992; Freer-Smith and Taylor, 1992;
27      Simini et al., 1992;  Taylor, Jr. and Hanson, 1992). This emphasis on foliar effects in the forest
28      effects literature (rather than marketable yield) is similar to the state of science for agricultural
29      crops prior to 1980. More recent studies address the effects of air pollutants on forest tree
30      species diversity (Bringmark and Bringmark, 1995; Vacek et al., 1999; Weiner et al., 1997).
31      However, such information is of limited use in economic assessments. The exception is in

        January 2005                             9-322        DRAFT-DO NOT QUOTE OR CITE

-------
 1      measuring the economic value of aesthetic changes in a forest stock, where changes in species
 2      composition may affect recreational values (Crocker, 1985).
 3           The data concerning changes in marketed output such as board-feet of lumber or changes
 4      in growth rates in managed forests, the affects on the growth of almond, peach, apricot, pear and
 5      plum trees in orchards cited in the 1996 document (U.S. Environmental Protection Agency,
 6      1996) have not been quantified.  In addition, the economic impact of reductions in growth of
 7      seedlings of evergreen trees, e.g., slash pine, presented in the same document have not been
 8      valued. The few studies which attempted to measure economic losses arising from exposures to
 9      O3 or other pollutants circumvented the lack of plant science data by assuming (often arbitrary)
10      reductions in forest species growth or harvest rates (Adams, 1986; Callaway et al., 1986;
11      Crocker and Forster, 1986; Haynes and Adams,  1992). Although the economic estimates
12      reported in the 1996 AQCD are comparable to those reported for agriculture (e.g., $2.6 billion
13      for eastern Canada forests, $2.9 billion for eastern U.S. forests  in 2000 dollars), the lack of yield
14      and/or growth effects data makes these studies only suggestive at best, of the economic
15      consequences of forest effects directly attributable to O3 exposures questionable. Recent
16      developments in forestry economic modeling capabilities, in support of climate change research,
17      have enhanced the  ability to measure the effects of environmental  stressors on this sector
18      (Adams et al.,1996; McCarl et al., 1998).  However, these models  need data on changes in either
19      timber production or growth rates, both of which are lacking for forest species under alternative
20      O3 levels.
21
22      9.8.4  Studies Since 1996 of Ozone Exposure Effects on the Economic Value of
23             Agriculture, Forests, and Ecosystems
24           Of the few current (post-1996) economic studies addressing  agricultural effects, none offer
25      new insights of value in determining the economic cost of O3 exposures. These post-1996
26      studies used variants of the economic methods from earlier assessments and measure yield
27      changes from response functions arising from the NCLAN or similar data.  For example, Kim
28      et al. (1998) used a mathematical programming model of the San Joaquin Valley agricultural
29      sector in California, combined with crop yield response functions, to assess the economic effects
30      of O3 on California crops. Their results showed net benefits from  reductions in ambient O3
31      levels, a finding  consistent with all previous economic  assessments.  In another study,

        January 2005                             9-323       DRAFT-DO NOT QUOTE OR CITE

-------
 1      Westenbarger and Frisvold (1994) measured agricultural exposures to O3 (and acid precipitation)
 2      in the United States. Though not an economic analyses of the costs of ambient exposures, they
 3      identified areas of the United States of greatest potential economic damage based on the
 4      interface between regional pollution levels and the value of crop production in each region.
 5           A study by Murphy et al. (1999) of the economic effects of tropospheric O3 on U.S.
 6      agriculture is of note here, because it confirms the general magnitude of economic effects
 7      reported by the two key studies performed a decade earlier (Kopp et al., 1985; Adams, 1986).
 8      Specifically, Murphy et al. (1999) evaluated benefits to eight major crops associated with several
 9      scenarios concerning the reduction or elimination of O3 precursor emissions from motor vehicles
10      in the United States. Their analysis reported a $2.8 to 5.8 billion (1990  dollars) benefit from
11      complete  elimination of O3 exposures from all sources, i.e., ambient O3 reduced to a background
12      level assumed to be 0.025 to 0.027 ppm. While the analytical framework is similar to Adams
13      et al. (1986) in the use of NCLAN-based yield response functions and a mathematical
14      programming-based economic optimization model, the study is novel in its focus on the role of
15      motor vehicle emissions of VOCs/NOx in total anthropogenic O3 levels. The  study is also
16      notable in its careful attention to federal farm program effects, particularly the deficiency
17      payment component.
18           In addition to these studies in peer-reviewed journals, a number of site-specific effects
19      studies have been performed, primarily by consulting companies for state public utility
20      commissions. Although perhaps of use to public utility commissioners concerned with effects
21      from single power plants or other localized sources, these regional studies generally contribute
22      little to the assessment of air pollution effects at the national level. Also, such reports are not
23      peer reviewed. Therefore, they are not discussed  here.
24           There have  been a number of recent studies  of air pollutant effects on tree species in the
25      literature.  Some have reported changes in total biomass and focused on European species
26      (Kurczynska et al., 1997).  Other studies have assessed changes in composition of forest species
27      (biodiversity) or forest health due to exposure to air pollutants (Bringmark and Bringmark, 1995;
28      McLaughlin and Percy,  1999; Vacek et al., 1999). As noted previously, changes in  forest
29      biomass and composition are more difficult to value than marketable products. However,
30      measures of forest composition or health have implications for an area of increasing policy
31      concern, that being the effect of air pollutants and other environmental stressors on unmanaged

        January 2005                              9-324       DRAFT-DO NOT QUOTE OR  CITE

-------
 1      (natural) ecosystems and the services they provide (Goulder and Kennedy, 1997; Pimentel et al.,
 2      1997). Considerable discussion has occurred among ecologists and economists as to the
 3      appropriate means for valuing these services (Anderson, 1990; Carpenter and Dixon, 1985;
 4      Common and Perrings, 1992).  A number of conceptual articles have been published on this
 5      issue in both economic and ecological journals (Bergstrom, 1990; Castle, 1993; Pearce, 1993;
 6      Suter, II, 1990).
 7           A continuing empirical challenge concerns the lack of information on how changes in
 8      biodiversity affect ecosystem performance resulting from O3 stresses, and the problem of
 9      establishing economic values for  such changes (Cairnes and Lackey, 1992; Norton,  1987; Pimm,
10      1997; Polasky, 2001; Randall, 1991). As noted in the 1996 AQCD, and more recently by Daily
11      (1999, 2000) and Polasky (2001), there continues to be a lack of empirical  studies that actually
12      assess the economic value of changes in biodiversity or in service flows due to any
13      environmental stressors. While some studies report monetary estimates, the estimates are
14      generally for expository purposes and those would not be as defensible as the agricultural studies
15      described earlier. For example, Costanza et al. (1997) assigned a value to the world's
16      ecosystems, but the procedures used render this an exploratory study at best. As assessed by
17      Polasky (2001), "In general, the field of valuation of ecosystem services is in its infancy."
18      He attributed the lack of empirical studies due to both a lack of the understanding of ecology of
19      ecosystem services and to the absence of reliable methods to estimate the value of these services.
20           In summary, the studies of crop and forest responses in the economic literature indicate
21      that O3 reduces crop yields and imposes economic costs.  The economic literature also indicates
22      that O3 adversely influences the physiological performance of tree species and demonstrates, as
23      expected, that  changes in growth have economic consequences. However, the economic data
24      and literature available on ecosystem effects of O3 exposures are not sufficient to determine the
25      economic costs.
26
27      9.8.5  Limitations of Scientific Studies and Economic Information
28           The 1996 O3 AQCD discussed the need for additional research on both ecological
29      functions and economic methodology in order to better understand the economic implications of
30      air pollutants on ecosystem services. As noted by Daily (1999, 2000), Polasky (2001), and
31      others, this research agenda continues to be important. Despite the large number of discussion

        January 2005                             9-325        DRAFT-DO NOT QUOTE OR CITE

-------
 1      and survey articles published since 1996, there still are not sufficient data by which to estimate
 2      confidently the magnitude of economic effects of O3 to forests and natural ecosystems. Nor is it
 3      apparent that ongoing research is adequate to answer this question in the next criteria document
 4      cycle. Specifically, there does not appear to be any comprehensive, ecological studies underway
 5      that attempt to measure changes in ecosystem outputs under alternative O3 or other air pollutant
 6      levels.  Thus, in the near term, ecosystem services can only be discussed in qualitative terms.
 7      However plausible the likelihood of economic damages to  ecosystems, the available scientific
 8      and economic information does not provide specific guidance on the magnitudes of these effects.
 9           Beside the need to improve our understanding of the effects of O3 exposures on natural
10      ecosystems, a number of areas of research could help assess the full economic consequences of
11      such pollutants on managed ecosystems.  The first of these is the relationship between O3
12      exposure levels and the variability of crop yields or changes in forest biodiversity. Most
13      assessments are based on the average or expected yield response of a crop to air pollution
14      exposure. However, the variability in yields (the spread or dispersion around the mean) appears
15      to be affected by the nature of plant exposure to pollutants  (Hogsett et al.,  1985; Musselman
16      et al., 1983). Plants exposed to the same mean dose but with different second moments of the
17      distribution of exposure may have different mean yields (Altshuller and Lefohn, 1996; Lefohn
18      and Benedict, 1982).  In addition, the variability of the yield of the plant may also be increased
19      by greater variability in exposures (e.g., a higher frequency of extreme events).  The economic
20      significance of higher yield varieties is such that variability may impose additional economic
21      costs, because most farmers have been reported to be averse to risk and prefer less variability for
22      a given yield (or profit). To date, no economic assessments of O3 damages to agriculture or
23      vegetation in general include risk-averse behavior (studies  cited here assumed risk neutrality).
24      To assess the economic consequences of a relationship between farmers' risk preferences and
25      O3-induced changes in yield variability will require more information on the potential effects of
26      changes in O3 on crop yield variability.  While no economic studies were found on the effects of
27      O3-induced changes in yield variability, a number of recent studies of climate change effects on
28      crop yields have indicated increased economic costs in the  presence of increased climatic
29      variability (e.g., Mearns et al., 1997; Dixon and Segerson,  1999).  Analogous economic costs
30      would be expected for changes in air pollution distributions; and these effects need to be
31      examined and quantified.

        January 2005                              9-326        DRAFT-DO NOT QUOTE OR CITE

-------
 1           Another research area concerns the need improvement in our understanding of temporal
 2      (dynamic) and spatial characteristics of O3 exposures and their implications for crop yields,
 3      production and producer profits.  Most economic studies are static, in that they compare two
 4      states of the world (e.g., economic activity at one O3 level versus at an alternative level).  In
 5      addition, most national- level studies (the type needed to evaluate SNAAQS) display coarse
 6      regional-level resolution in terms of crop response, O3 exposure, and economic behavior. The
 7      responses of producers to changes in yields due to changes in O3 levels are generally assumed to
 8      be similar over geographical areas up to several states in size.  However, the changes between air
 9      quality scenarios are more likely to be characterized by transient changes in exposure levels,
10      which means the producer responses are also likely to be gradual, rather than abrupt. Similarly,
11      the lack of finer  scale (regional-level) data and modeling capabilities suggests that important
12      micro-level physical and economic effects are ignored.  To what extent these abstractions
13      influence net economic effects is an empirical question. Research on these types of abstractions
14      and assumptions within other economic settings, such as climatic change, have shown that they
15      have  implications for economic measurements (Adams, 2002).
16           Another issue is the natural or background level of O3 (or other pollutants of interest)
17      assumed in economic studies. While many economic studies focus on changes in pollution
18      levels from current conditions, some studies have measured the economic damages between an
19      assumed,  or pristine, level and current levels in  agricultural areas.  Such an analysis, it is
20      reasoned, will provide a measure of the net damages due to anthropogenic sources. The
21      challenge here is to have a correct measure of the background level of the pollutant. Recent
22      research by Lefohn et al. (2001) has suggested that background levels may be considerably
23      higher than assumed in some of the previous economic assessments reported in the 1996 AQCD
24      (25 to 30 ppb in  most studies). For example, Lefohn et al. (2001) detected hourly readings of
25      from  40 to 80 ppb during winter and spring months in remote areas of the United States.
26      If background O3 levels are in the range, then the economic damages estimated in studies with
27      lower background levels will be overstated.  The issue of the background O3 level  is important to
28      all assessments of vegetative damages due to O3.
29           In terms of expanding economic methods  for future assessment, analysts should consider
30      using more "reduced form" estimation methods, particularly in situations where the availability
31      of dose-response functions is limited. This estimation approach is exemplified by  Garcia et al.

        January 2005                              9-327       DRAFT-DO NOT QUOTE OR CITE

-------
 1      (1986). Specifically, their econometric study of the impact of O3 on producer profits used such a
 2      reduced-form approach. In this approach, farmer actions are modeled as a function of ambient
 3      O3 levels (along with other explanatory variables) without the direct use of dose response
 4      functions.  The advantage of this procedure is that one source of modeling uncertainty, the need
 5      for dose-response functions, including time-consuming crop experiments to generate data, is
 6      reduced. Actual responses of farmers' profits across air pollution gradients of ambient pollution
 7      are observed instead of hypothesized.  Although this procedure has not been widely used in air
 8      pollution economic assessments, it has been used in a number of relatively recent climate change
 9      studies (e.g., Mendelsohn et al. [1994]). The reduced-form method suffers from the fact that if
10      proposed O3 levels are lower than those observed in the estimation sample, then the prediction
11      accuracy of the method deteriorates. Also, some dose-response information is needed, if only to
12      establish the plausibility of the economic estimates.
13           Another area that may improve economic assessments is incorporating consideration of
14      livestock issues. To date, most agricultural economic assessments of O3 impacts have ignored
15      the livestock sector. Presumably this is because ambient O3 levels do not noticeably affect meat
16      yields. However, if feed prices and pasture conditions are affected by ambient O3 levels, then a
17      more accurate estimate of economic impacts would be forthcoming by including this link to
18      livestock in the assessment. These types of feed production and feed price effects are included
19      in the mathematical programming  model used in Adams et al. (1986), but not in most other O3
20      effects assessments. The significance of livestock-feed linkages are demonstrated in a recent
21      study in the Netherlands by Kuik et al (2000).  Using a mathematical programming model
22      similar to that in Adams et al. (1986), Kuik et al. (2000) found that livestock effects were
23      prominent, due mainly to improved pasture yields under reduced ambient O3 levels.
24
25      9.8.6  Conclusions
26           Substantial progress has been made over the past two decades in our understanding of the
27      effects of O3 and other oxidants on vegetation, particularly for agriculturally important plant
28      species. The physical and economic effects on agriculture are well documented and provide
29      useful information for the setting of SNAAQS.  Effects on forests and natural  ecosystems remain
30      problematic, due to limitations in natural science data and economic methods. The problem is
31      most acute for valuing natural ecosystem service flows.

        January 2005                              9-328       DRAFT-DO NOT QUOTE OR CITE

-------
 1           The current limitations surrounding forests and natural ecosystems present a rich research
 2      agenda. However, not all research needs are likely to lead to better policies.  Thus, areas of
 3      greatest potential value in terms of regional policymaking need to be prioritized.  Such priority
 4      setting can be assisted by sensitivity analyses with existing economic models. By measuring the
 5      changes in economic effects arising from changes in key parameters, research data gaps most
 6      likely to affect economic values can be identified.
 7
 8
 9      9.9  SUMMARY AND CONCLUSIONS FOR VEGETATION AND
10           ECOSYSTEM EFFECTS
11      9.9.1 Introduction
12           A significant number of ozone-related studies  were published between  1996 and 2004, and
13      they are reviewed in this document in the context of earlier publications reviewed in the previous
14      criteria documents (U.S. Environmental Protection Agency, 1878;  1986; 1992; 1996).
15      In general, there has been a shift away from chamber studies in favor of more field-based
16      approaches although chamber exposures still dominate the effects literature.  Field-based
17      approaches include increased survey's of visible injury as well as free-air exposure systems that
18      eliminate  some of the problems associated with chamber effects. Increased emphasis has also
19      been placed on quantifying aspects of ozone uptake to better link ambient exposure monitoring
20      with plant/tree response. Much of the progress in quantifying uptake has occurred in Europe
21      with application of models simulating uptake.  Evaluation of this new information has added to
22      our knowledge, but has not fundamentally  altered the conclusions of previous Ozone Criteria
23      Document (U.S. Environmental Protection Agency, 1996).
24           It is known that ozone is phytotoxic,  and that toxicity occurs  only if ozone or its reaction
25      products reach the target tissues in the plant cell. Recent studies have provided an increased
26      understanding of how ozone interacts with the plant at the cellular  level. Progress has been
27      made on understanding the initial stages of ozone toxicity, which resemble a "wound" response
28      in plants.  The alteration of normal metabolism due  to a "wound" starts in the cytoplasm, and
29      'spreads' to other cellular organelles. One of the secondary reactions is linked to an activation of
30      a senescence response that results in loss of the critical enzymatic components of photosynthesis.
31      The loss of photosynthetic capacity is in turn linked to the lowered productivity of plants and

        January 2005                             9-329       DRAFT-DO NOT QUOTE OR CITE

-------
 1      problems with efficient translocation of carbon. Although photosynthesis and translocation still
 2      occurs at a reasonable rate, the loss of productivity, while better understood than in 1996, is still
 3      not clearly explained. Dramatic strides in understanding the genetic make-up of plants, gene
 4      control, and signal transduction/control over the last few years will accelerate in the future.  This
 5      understanding will translate into better models, more detailed schemes of how O3 alters much of
 6      the basic metabolism of plants, and how to construct an index that better quantifies exposure and
 7      effect. However, the translation of those mechanisms into how O3 is involved with the altered
 8      cell metabolism and subsequent reductions in whole plant productivity and other physiological
 9      facts has not yet been fully solved.
10           A number of biotic and abiotic factors are known, or suspected, to alter plant response to
11      ozone. New information published since 1996 does not significantly change the conclusions
12      arrived at in the previous criteria document regarding these multiple interactions.  A major
13      concern from  the last review was the potential, but unquantified, interaction of ozone and
14      diseases and insects.  The nature of these interactions with O3 are probably dictated by the
15      unique, highly specific biochemical relationships between pathogen and host plant. At this time,
16      therefore, although some diseases may become more widespread or severe as a result of
17      exposure to O3, it is still not possible to predict which diseases are likely to present the greatest
18      risks to crops  and forests.
19           Considerable  emphasis during the last decade has been placed on research into O3
20      interactions with the components of global climate change: increased atmospheric CO2,
21      increased mean global temperatures, and increased surface level UV-B radiation.  To date, the
22      limited experimental evidence and that obtained by computer simulation suggest that even
23      though an enriched CO2 atmosphere (-600 ppm) would more than offset the impact of O3 on
24      responses as varied as wheat yield or young Ponderosa growth, the concurrent increase in
25      temperature would reduce but probably not eliminate the net gain. A similar decrease in the net
26      gain resulting from the complete reversal by  CO2 of the inhibition of photosynthesis caused by
27      O3 has been reported for increased UV-B irradiation. Overall, results are preliminary, based
28      upon minimal data, and do not allow adequate prediction of ozone effects under future climate
29      scenarios. Temperature is unquestionably also an important variable affecting plant response to
30      O3 in the presence of the elevated CO2 levels contributing to global climate change. In contrast,
31      evidence continues to accumulate to indicate that exposure to O3 sensitizes plants to low

        January 2005                               9-330        DRAFT-DO NOT QUOTE OR CITE

-------
 1      temperature stress, by reducing belowground carbohydrate reserves, which may lead to
 2      responses in perennial species ranging from rapid demise to impaired growth in subsequent
 3      seasons.
 4           Efforts increased in the area of relating ozone flux (e.g., uptake) to response. Flux is
 5      affected by a host of biotic and abiotic factors, including boundary layer conditions, stomatal
 6      conductance, moisture conditions, solar irradiance and others. All of these factors provide a
 7      more accurate estimation of flux for predicting plant, crop, or forest response. While this area of
 8      research is promising, currently there is insufficient data linking flux with response across a
 9      range of species or in different ecosystems. In addition, a flux-based standard would require
10      selection of a particular species or species assemblage across a meteorologically similar area in
11      order to adequately model the range of factors known to influence uptake.  At this time, there is
12      insufficient information available to employ this approach nationally, but with continued
13      research may eventually provide an alternative index to the current exposure index.
14
15      9.9.2  Methodology
16           The majority of ozone effects studies are fumigation studies conducted in controlled
17      chambers, as noted in the previous document. The previous document noted that open-top
18      chambers (OTC's) represent the best technology for determination of crop yield to ozone at the
19      present time. While OTC's are still the best method for conducting controlled exposures of
20      varying length and frequency for exposure-response relationships, several new approaches have
21      been applied to ozone research, most notably free-air exposure or FACE systems.  FACE
22      systems eliminate many of the concerns raised about closed or open-top chamber experiments,
23      including small plot size, altered microclimate within OTC's, and the effect of charcoal filtering
24      on overall air quality within OTC's.  Although FACE systems have increased our understanding
25      in some areas, in most cases results from FACE systems have confirmed what we already
26      understood or hypothesized about how plants and plant assemblages respond to ozone.
27      In particular, studies with aspen showed that ozone symptom expression was generally similar in
28      OTC's, FACE, and also sites along an ambient ozone gradient, supporting the previously
29      observed variation among aspen clones using OTC's (Isebrands et al., 2000; 2001; Karnosky
30      et al., 1999).  In addition, root growth is often found to be the most sensitive biomass response to
31      ozone, as reported in the previous document.  Caution is still needed, however when making

        January 2005                             9-331        DRAFT-DO NOT QUOTE OR CITE

-------
 1      quantitative extrapolations from OTC's, and free-air exposure systems will continue to provide
 2      valuable information in support of scaling efforts.
 3           The lack of rural monitors continues to be a major problem in the characterization of ozone
 4      exposures in remote areas, and in linking effects with exposure in natural ecosystems.  Since the
 5      last document, the use of passive samplers have expanded monitoring efforts to remote areas that
 6      were previously uncharacterized.  This has greatly enhanced our ability to link ozone
 7      symptomology with elevated exposure in remote areas.  Passive samplers do not capture
 8      temporal dynamics of exposure, however; therefore, passive samplers are not a substitute for
 9      active monitors when attempting to link dynamics of exposure with plant response, or when
10      developing exposure/dose response relationships that are needed in the standard setting process.
11           Exclusion methods such as EDU are the least disruptive of ambient culture conditions in
12      the field, as noted in the previous document. However, the level of protection afforded by EDU
13      is site and species specific, and  is subject to local meteorologic conditions. Because of the
14      variability observed in the level of protection provided, and the fact that mechanisms of
15      protection afforded by EDU and other exclusion methods are unknown, caution is needed in
16      applying this approach to the study of ozone effects in the field.
17
18      9.9.3  Mode-of-Action
19           There are several steps in  the process of ozone uptake and toxicity that are better
20      understood than in 1996 and are important in developing more refined hypotheses on ozone
21      uptake and mode of action on plants, and developing a flux-based index for use in quantifying
22      response, and ultimately are relevant for developing an air quality standard. Theses are listed
23      below:
24        (1)  The entrance of O3 into  the leaf through the stomata remains the critical step in O3
               sensitivity.  Not only does O3 modify the opening of the stomata, usually closing it
               partially, but O3 also appears to alter the response of stomata to other stressful
               situations, including a lowering of water potential and ABA responses. The
               concentration of O3 within the leaf is not the same as the external concentration due
               to reactions within the leaf but it is not "zero" (Moldau and Bichele, 2002).
25        (2)  The initial reactions of O3 within the leaf are still unclear but the involvement of
               hydrogen peroxide is clearly indicated. The detection of possible products by EPR
               has progressed but is still not at the point where  any products can be identified.
               Nonetheless, reaction of O3 or its product with ascorbate and possibly other

        January 2005                             9-332       DRAFT-DO NOT QUOTE OR CITE

-------
               antioxidants present in the apoplastic space of the mesophyll cells is clear and serves to
               lower the amount of O3 or product available to alter the plasma membrane of the cells.

26        (3)   The initial sites of membrane reactions seems to be at the transport properties and
               possibly the external signal transducer molecules. The alteration and mechanism of
               the alteration of the varied carriers of K+and Ca2+ is far from clear but it would seem
               that one of the primary triggers of O3-induced cell responses is due to a change in
               internal Ca2+ levels (Cazale et al., 1998).

27        (4)   The primary set of metabolic reactions that O3 triggers is now clearly the "wounding"
               responses like that generated by cutting of the leaf or by pathogen/insect attack.  This
               seems to be due to a rise in cytoplasmic Ca2+. Ethylene release and alteration of
               peroxidases  and PAL activities as well as activation of many "wound"-derived genes
               seem to be linked and some of the primary reactions to ozone exposure.

28        (5)   The alteration of normal metabolism due to a "wound" has affects outside of the
               cytoplasm.  What effects are due to the "spreading of the problem" to other cellular
               organelles is less clear.  One of the secondary reactions is linked to an activation of a
               senescence response. Certainly the loss of Rubisco and its messenger RNA is linked to
               an early senescence or a speeding up of normal  development leading to senescence.
               Certainly the loss of photosynthetic capacity is linked to the lowered productivity of
               plants and problems with efficient translocation are indicated. Although photosynthesis
               and translocation still occurs in a reasonable rate, the loss of productivity is not
               clearly explained.

29           The new information available  on the mode of action of ozone is in part a result of

30      improved molecular tools for following rapid changes that occur within the leaf.  Clearly there

31      are many changes that occur within hours or possibly days of the exposure to O3. However,

32      there is another effect due to O3 which takes longer to occur and tends to be most obvious under

33      exposure to low O3 concentrations for long periods.  These effects have been linked to

34      senescence or some  physiological response very closely linked to senescence (i.e., translocation,

35      reabsorption, allocation of nutrients and carbon).

36           Langebartels et al.  (1997) have discussed "memory" or "carry-over effects" within the

37      plant's to explain sensitivity to  frost in the winter after exposure to ozone in the summer. Others

38      have argued that it is nutrient status of the tree during the over-wintering phase of its life and

39      chronic exposure to  ambient O3 (less severe with fewer peaks of very high levels) which induces

40      (1) mineral nutrient  deficiency; (2) alterations of normal metabolism, including translocation and

41      allocation of carbohydrates and probably nitrogen; (3) disturbance of normal transpiration and

42      diurnal cycling, leading to water stress (Schmieden and Wild, 1995). While general nutrient

43      concentrations within the foliage may not occur, it may be that localized deficiencies do. This is


        January 2005                             9-333        DRAFT-DO NOT  QUOTE OR CITE

-------
 1      hard to observe or prove without a great deal of work on all portions of a tree and without a
 2      general hypothesis of what is occurring.
 3           It is important to note that the dramatic strides in understanding the genetic make-up of
 4      plants, gene control, and signal transduction/control over the last few years will only accelerate
 5      in the future.  That understanding will translate into better models of the hypotheses listed and
 6      more detailed schemes of how O3 alters much of the basic metabolism of plants.  Thus overall,
 7      understanding of how O3 interacts with the plant at a cellular level has been likewise
 8      dramatically improved. However, the translation of those mechanisms into how O3 is involved
 9      with the altered cell metabolism and subsequent reductions in whole plant productivity and other
10      physiological  facts has not yet been fully solved. As the understanding of wounding responses
11      of plants and more genome details and varied plant mutants become available, the cellular and
12      physiological  responses of plants  to O3 exposures is slowly becoming clearer. However, more
13      studies need to be  made on a larger variety of species before this information can be
14      incorporated into indices of response and a protective standard.
15
16      9.9.4  Modification of Growth Response
17           There have been  only a few studies reported since the 1996 document and none that
18      significantly change the conclusions of the previous criteria document with regard to factors
19      known to alter plant response to ozone.
20           In the area of biotic interactions, new evidence with regard to insect pests and diseases has
21      done little to remove the uncertainties noted in the previous criteria document (see Docherty
22      et al.,  1997; and Fluckiger et al., 2002 for recent reviews).  Most of the large number of such
23      interactions that may affect crops, forest trees and other natural vegetation have yet to be studied.
24      The trend suggested previously that O3 increases the likelihood and success of insect attack has
25      received some support from recent studies, but only with respect to chewing insects. With the
26      economically  important group of  sucking insects such as the aphids, no clear trends have been
27      revealed by the latest studies.  Hence, although it seems likely that some insect problems could
28      increase as a result of increased O3 levels, we are still far from being able to predict the nature of
29      any particular O3-plant-insect interaction, its likelihood or its severity.
30           The situation is a little clearer with respect to interactions involving facultative,
31      necrotrophic plant pathogens, with O3 generally leading to increased disease. With obligate,

        January 2005                              9-334       DRAFT-DO NOT QUOTE OR CITE

-------
 1      biotrophic fungal, bacterial and nematode diseases, however, there are twice as many reports
 2      indicating O3-induced inhibitions than enhancements.  The frequent reports that infection by
 3      obligate biotrophs reduces the severity of O3-induced foliar injury does not result in true
 4      "protection" since there are negative effects on the host plant of the disease/?er se.  With
 5      obligate biotrophs, the nature of any interaction with O3 is probably dictated by the unique,
 6      highly specific biochemical relationships between pathogen and host plant. At this time,
 7      therefore, although some diseases may become more widespread or severe as a result of
 8      exposure to O3, it is still not possible to predict which diseases are likely to present the greatest
 9      risks to crops and forests.
10           Recent studies have done little to improve our understanding of the nature of interactions
11      between O3 and root symbionts.  Several studies have indicated that the functioning of tree root
12      symbioses with mycorrhizae may be adversely affected by O3, but there is also evidence that the
13      presence of mycorrhizae may overcome root diseases stimulated by O3.  There is also evidence
14      that O3 may encourage the spread of mycorrhizae to the roots of uninfected trees.
15           The few recent studies of the impact of O3 on intra-specific plant competition confirmed
16      that grasses frequently show greater resilience than other types of plants.  In grass-legume
17      pastures, the leguminous species suffer greater growth inhibition (Johnson et al., 1996a;
18      Nussbaum et al.,  2000a).  And the suppression of Ponderosa pine seedling growth by blue wild-
19      rye grass was markedly increased by O3 (Andersen et al., 2001). However, we are far from
20      being able to predict the outcome of the impact of O3 on specific competitive situations such as
21      success!onal plant communities, or crop-weed interactions.
22           Although some recent field studies have indicated that O3 impact significantly increases
23      with increased ambient temperature, other studies have revealed little effect of temperature.
24      Light, a  component of the plant's physical environment, is an essential 'resource' whose energy
25      content drives photosynthesis and CO2 assimilation. It has been suggested that increased light
26      intensity may increase the sensitivity to O3 of light-tolerant species, but decrease that of shade-
27      tolerant  species, but this appears to be an over-simplification with many exceptions.
28      Temperature affects the rates of all physiological processes based on enzyme-catalysis and
29      diffusion, and each process and overall growth (the integral of all processes) has a distinct
30      optimal temperature range. It is important to note that a plants response to changes in
31      temperature will  depend on whether it is growing near its optimum temperature  for growth or

        January  2005                             9-335        DRAFT-DO NOT QUOTE OR CITE

-------
 1      near its maximum temperature (Rowland-Bamford, 2000).  But temperature is unquestionably an
 2      important variable affecting plant response to O3 in the presence of the elevated CO2 levels
 3      contributing to global climate change. In contrast, evidence continues to accumulate to indicate
 4      that exposure to O3 sensitizes plants to low temperature stress, by reducing belowground
 5      carbohydrate reserves, which may lead to responses in perennial species ranging from rapid
 6      demise to impaired growth in subsequent seasons (i.e., 'carry-over effects').
 7           Although the relative humidity of the ambient air has generally been found to increase the
 8      adverse effects of O3 by increasing stomatal conductance and thereby increasing O3 flux, there is
 9      abundant evidence that the ready availability of soil moisture results in greater sensitivity to O3.
10      The partial 'protection' against the adverse effects of O3 afforded by drought has been observed
11      in field experiments and modeled in computer simulations (Broadmeadow and Jackson, 2000).
12      There is also compelling evidence that O3 can predispose plants to drought stress (Maier-
13      Maercker,  1998).  Hence the response will depend to some extent upon the sequence in which
14      the stresses occur, but, even though the nature of the responses is largely species-specific,
15      successful  applications of model simulations have shown the way to larger scale predictions of
16      the consequences of O3 x drought interactions. However, it must be recognized that regardless
17      of the interaction, the net result on growth in the short-term is negative, although in the case of
18      tree species, other responses such as increased water use efficiency could be a benefit to survival
19      in the long term.
20           Somewhat analogously with temperature, it appears that any shift away from the nutritional
21      optimum may lead to greater ozone sensitivity, but the shift would have  to be substantial before
22      a significant effect on response to O3 was observed. Mineral nutrients in the soil, other gaseous
23      air pollutants, and agricultural chemicals constitute chemical factors in the environment. The
24      evidence regarding interactions with specific nutrients is still contradictory: there is some
25      experimental evidence that low general fertility increases sensitivity to O3, while simulation
26      modeling of trees suggests that nutrient deficiency and O3 act less than additively, but there are
27      too many example of contrary trends to permit any sweeping conclusions.
28           Interactions of O3 with other air pollutants have received relatively little recent attention
29      since 1996 (see Barnes and Wellburn, 1998; and Fangmeier et al., 2002 for recent reviews).
30      The situation with SO2 remains inconsistent, but seems unlikely to pose  any additional risk to
31      those related to the individual pollutants. With the NO and NO2 the situation is complicated by

        January 2005                               9-336        DRAFT-DO NOT QUOTE OR CITE

-------
 1      their nutritional value as a source of N.  In leguminous species, it appears that NO2 may reduce
 2      the impact of O3 on growth, with the reverse in other species, but the nature of the exposure
 3      pattern, i.e., sequential or concurrent, also determines the outcome. Much more investigation is
 4      needed before we will be able to predict the outcomes of different O3-NO-NO2 scenarios.  The
 5      latest research into O3 * acid rain interactions has confirmed that, at realistic acidities, significant
 6      interactions are unlikely. A continuing  lack of information precludes offering any
 7      generalizations about interactive effects of O3 with NH3, HF, or heavy metals. More evidence
 8      has been reported for protective effects  against O3 afforded by the application of fungicides.
 9           Considerable emphasis during the last decade has been placed on research into O3
10      interactions with the components of global climate change: increased atmospheric CO2,
11      increased mean global temperatures, and increased surface level UV-B radiation. However, it
12      must be noted that most of these studies have tended to regard increased CO2 levels and
13      increased mean temperatures as unrelated phenomena.  Experiments into the effects of doubled
14      CO2 levels at today's mean ambient temperatures are of questionable value in trying to assess the
15      impact of climate change on responses to O3. To date, the limited experimental evidence and
16      that obtained by computer simulation suggest that even though an enriched CO2 atmosphere
17      (-600 ppm) would more than offset the impact of O3 on responses as varied as wheat yield or
18      young Ponderosa growth, the concurrent increase in temperature would reduce but probably not
19      eliminate the net gain.  A similar decrease in the net gain resulting from the complete reversal by
20      CO2 of the inhibition of photosynthesis  caused by O3 has been reported for increased UV-B
21      irradiation.  However, these are preliminary results based upon minimal data.
22
23      9.9.5 Exposure Indices
24           The previous Criteria Document (U.S. Environmental Protection Agency,  1996a) noted
25      that the complexities associated with uptake and ozone interactions with external physical and
26      internal genetic factors that influence plant response makes the development of an 'ideal'
27      exposure index that characterizes plant exposure and response extremely difficult. As a result, a
28      biologically relevant surrogate for uptake was recommended that related ambient exposure to
29      measured growth/yield response. Despite additional research linking estimates of flux with plant
30      response since 1996, there is still insufficient information to identify a flux-based model that
31      incorporates sufficient complexity across space and time to be non-site or non-species specific.

        January 2005                            9-337       DRAFT-DO NOT QUOTE OR CITE

-------
 1      Therefore, based on the current state of knowledge, exposure indices that cumulate and
 2      differentially weight the higher hourly average concentrations, but include the mid-level values,
 3      still represents the best approach in the United States for relating vegetation effects to exposure.
 4           The few studies published in this interim have substantiated earlier conclusions on the role
 5      of exposure components, including concentration, duration and exposure patterns, in determining
 6      the growth response of plants to ozone (Yun and Laurence, 1999; Oksanen and Holopainen,
 7      2001).  Recent studies using different exposure patterns have confirmed earlier studies on the
 8      role of higher concentrations and exposure duration (Nussbaum et. al., 1995).  A role for peak
 9      concentrations is inferred based on improved air quality in regions like the San Bernardino
10      Mountains in southern California. The reductions of O3 in the San Bernardino area are
11      associated with reductions in the higher hourly average concentrations, but not in the midrange
12      hourly average concentrations, which are increasing (Lee et al., 2003; Lefohn and Shadwick,
13      2000).  General forest improvement is reported following a decrease of O3 along a decreasing
14      gradient of exposure (Miller and Rechel, 1999; Arbaugh et al., 2003; Tingey et al., 2004).  These
15      studies provide the basis for focusing on higher O3 concentrations, while including the lower
16      levels, when estimating the effects of emission reductions on vegetation.
17           New studies have  demonstrated the potential disconnection of peak events and maximal
18      stomatal conductance at xeric to mesic sites in California (Panek et al., 2002; Panek, 2004;
19      Grulke et al., 2002). In addition, a few studies have indicated the uptake of ozone during
20      nighttime hours is greater than previously thought (Grulke et al., 2003; Massman, 2004), and a
21      review of the literature suggests a large number of species exhibit some degree of conductance at
22      night (Musselman and Minnick, 2000). These studies suggest a reconsideration of the need to
23      cumulate exposure 24 hrs per day and not just during daylight  hours in exposure index.  This
24      lack of coincidence in temporal patterns of conductance and peak ambient concentrations
25      introduces uncertainty in assessing ozone's impact. The use of an exposure index that does not
26      consider regionally unique climate and site factors which modify stomatal conductance may, as a
27      result, under or over estimate growth effects.  The shortcomings of an ambient exposure-based
28      index is especially apparent when assessing ozone's potential impact across broad climatic
29      regions of the U.S. or Europe.
30           It is apparent that  additional research is needed to develop indices which are more
31      physiologically- and meteorologically-connected to the actual  dose of ozone the plant receives.

        January 2005                              9-338       DRAFT-DO NOT QUOTE OR CITE

-------
 1      The cumulative, concentration-weighted exposure indices are acknowledged surrogates for
 2      effective dose that are conceptually simple and easy to measure. As discussed in the previous
 3      Criteria Document, they do not fully characterize the potential for plant uptake and subsequent
 4      response since they do not include the physical, biological, and meteorological processes
 5      controlling the transfer of O3 from the atmosphere through the leaf and into the leaf interior.
 6           The flux-based approach should provide an opportunity to improve upon the
 7      concentration-based (i.e., exposure indices) approach.  A great deal of progress has occurred in
 8      development and testing of stomatal models that may be generally applicable across certain
 9      vegetation types (e.g., Matyssek et al., 2004; Grunhage and Jager 2003; Danielsson et al., 2003;
10      Emberson et al., 2000; Pleijel et al., 2000).  While a flux-based approach is preferred, a
11      cautionary argument was advanced in a few publications based on the non-linear relationship
12      between ozone uptake and foliar injury (growth not assessed). The concern is that not all O3
13      stomatal uptake results in  a reduction in yield, which depends to some degree on the amount of
14      internal detoxification occurring with each particular species.  Those species having high
15      amounts of detoxification potential may show little relationship between O3 stomatal uptake and
16      plant response (Musselman and Massman, 1999).
17           The European approach and acceptance of flux-based critical values is in part a recognition
18      of the problems associated with ambient exposure-based indices. Research continues across
19      Europe to develop the necessary experimental database and modeling tools that will be required
20      to provide the scientific basis for a critical level for ozone (Grunhage et al., 2004; Fuhrer et al.,
21      1997; Grunhage and Jager, 1994).
22           Given the current state of knowledge and the best available data, exposure indices that
23      cumulate and differentially weight the higher hourly average concentrations, but include the
24      mid-level values, continue to offer the most defensible approach for vegetation protection.
25      A large database exists that has been used for establishing exposure-response relationships, and
26      at this time, such a database does not exist for relating  ozone flux to growth response.
27           It is anticipated that  as the overlapping relationships of conductance, concentration, and
28      defense mechanisms are better defined, the flux-based  indices may be able to predict vegetation
29      injury and/or damage with more accuracy than the exposure-response models. However, it is
30      unclear that such is the case at this time.  The translation of these research and assessment tools
        January 2005                             9-339        DRAFT-DO NOT QUOTE OR CITE

-------
 1     to air quality standards has the additional need to be simple, understandable, and adaptive to a
 2     manageable monitoring program.
 3
 4     9.9.6 Ozone Exposure-Plant Response Relationships
 5           Data published since 1996 continue to support the conclusions of previous criteria
 6     documents that there is strong evidence that ambient concentrations of O3 cause foliar injury and
 7     growth and yield damage to numerous common and economically valuable plant and tree
 8     species. For annual vegetation, the data summarized in Table 4-19 show a range of growth and
 9     yield responses both within species and among species. Nearly all of these data were derived
10     from studies in OTCs, with only two studies using open-air systems in the UK (Ollerenshaw
11     et al., 1999;  Ollerenshaw and Lyons, 1999). It continues to be difficult to compare studies that
12     report O3 exposure using different indices,  such as AOT40, SUM06, or 7-h or 12-h mean values.
13     However, when such comparisons can be made, the results of recent research confirm earlier
14     results summarized in the previous criteria document (U.S. Environmental Protection Agency,
15     1996). A summary of earlier literature concluded that a 7-h 3-month mean of 49 ppb
16     corresponding to a SUM06 exposure of 26 ppm-h, would cause 10% loss in 50% of 49
17     experimental cases (Tingey et al., 1991). Recent data summarized in Table 4-19 support this
18     conclusion, and more generally indicate that ambient ozone exposures can reduce the growth and
19     yield of annual species. Some annual species such as soybean are more sensitive, and greater
20     losses may be  expected (Table 4-19). Thus the recent  scientific literature supports the
21     conclusions  of the previous criteria document that ambient O3 concentrations are reducing the
22     yield of major crops in the United States.
23           Much research in Europe has used the AOT40 exposure statistic, and substantial effort has
24     gone into developing Level-1 critical levels for vegetation using this index. Based on regression
25     analysis of 15  OTC studies of spring wheat including one study from the U.S. and 14 from
26     locations ranging from southern Sweden to Switzerland, an AOT40 value of 5.7 ppm-h was
27     found to correspond to a 10% yield loss, and a value of 2.8 ppm-h corresponded to a 5% yield
28     loss (Fuhrer et al., 1997).  Because a 4 to 5% decrease could be detected with a confidence level
29     of 99%, a critical level of an AOT40 value of 3 ppm-h was selected in 1996 (Karenlampi and
30     Skarby, 1996).
       January 2005                             9-340       DRAFT-DO NOT QUOTE OR CITE

-------
 1           In addition to reductions in crop yield, O3 may also reduce the quality or nutritive value of
 2      annual species. Many studies have shown effects of O3 on various measures of plant organs that
 3      affect quality, with most studies focusing on characteristics important for food or fodder. These
 4      studies indicate that there may be economically important effects of ambient O3 on the quality of
 5      crop and forage species. Previous criteria documents have concluded that visible symptoms on
 6      marketable portions of crops and ornamental plants can occur with seasonal 7-h mean O3
 7      exposures of 40 to 100 ppb (U.S. Environmental Protection Agency,  1978; 1986; 1996).  The
 8      recent scientific literature does not refute this conclusion.
 9           The use of OTCs may reverse the usual vertical gradient in O3 that occurs within a few
10      meters above the ground surface (Section 4.2).  This reversal suggests that OTC studies may
11      overestimate to some degree the effects of an O3 concentration measured several meters above
12      the ground.  However such considerations do not invalidate the conclusion of the previous
13      criteria document that ambient O3 exposures are sufficient to reduce  the yield of major crops in
14      the United States.
15           As for single-season agricultural crops, yields of multiple-year  forage crops are reduced at
16      ozone exposures that occur over large areas of the U.S. This result is similar to that reported in
17      the previous criteria document (U.S. Environmental Protection Agency, 1996). When  species
18      are grown in mixtures, O3 exposure can increase the growth of tolerant species while
19      exacerbating the growth decrease of O3-sensitive species. Because of this competitive
20      interaction, the total growth of the mixed-species community may not be affected by O3
21      exposure. However, in some cases mixtures of grasses and clover  species have shown
22      significant decreases in total biomass growth in response to O3 exposure in studies in the U.S.
23      and in Sweden. In Europe, a provisional critical level for herbaceous perennials of an AOT40
24      value of 7 ppm-h over six months has been proposed to protect sensitive plant species from
25      adverse effects of O3.
26           For deciduous tree species, recent evidence from free air exposure systems and OTCs
27      supports results observed previously in OTC studies.  For example, a series of studies was
28      undertaken using free air O3 enrichment in Rhinelander, WI (Isebrands et al., 2000; 2001).
29      These studies showed that O3 symptom expression was generally similar in OTCs, FACE, and
30      also sites along an ambient O3 gradient, supporting the previously observed variation among
        January 2005                              9-341        DRAFT-DO NOT QUOTE OR CITE

-------
 1      aspen clones using OTCs (Karnosky et al., 1999). As has been observed in previous criteria
 2      documents, root growth often is found to be the most sensitive biomass response to O3.
 3           Results since 1996 support the conclusion of the previous criteria document (U.S.
 4      Environmental Protection Agency, 1996) that deciduous trees are generally less sensitive to
 5      O3 than are most annual plants, with the exception of a few very sensitive genera such as
 6      Populus and sensitive species such as black cherry.  However, the data presented in Table 4-20
 7      suggest that ambient exposures that occur in the U.S. can sometimes reduce the growth of
 8      seedlings of deciduous species.  Results from multiple-year studies sometimes show a pattern of
 9      increased effects in subsequent years.  In some cases however, growth decreases due to O3 may
10      become less significant or even disappear over time. While some mature trees show greater O3
11      sensitivity than do seedlings in physiological parameters such as net photosynthetic rate, these
12      effects may not translate into measurable reductions in biomass growth.  However, because even
13      multiple-year experiments do not expose trees to O3 for more than a small fraction of their life
14      span, and because competition may in some cases exacerbate the effects of O3 on individual
15      species, determining effects on mature trees remains a significant challenge.
16           In Europe, a Level I critical level has been set for forest  trees based on OTC studies of
17      European beech seedlings.  A critical level was defined as  an AOT40 value of 10 ppm-h for
18      daylight hours for a 6-month growing season (Karenlampi  and Skarby, 1996). However, other
19      studies show that other species such as silver birch may be more sensitive to O3 than is beech
20      (Paakkonen et al., 1996).
21           For evergreen tree species as for other tree species, the O3 sensitivity of different genotypes
22      and different species varies widely.  Based on studies with seedlings in OTCs, major species in
23      the US are generally less sensitive than are most deciduous trees, and slower growing species are
24      less sensitive than are faster growing species. There is evidence that interacting stresses  such as
25      competition may increase the sensitivity of trees to O3.  As for deciduous species, most
26      experiments with evergreen species have only covered a very  small portion of the life span of a
27      tree, and have been conducted with seedlings, so estimating effects on mature trees is difficult.
28           For all types of perennial vegetation, cumulative effects  over more than one growing
29      season may be important, and studies for only a single season  may underestimate effects.
30      Mature trees may be more or less sensitive to O3 than are seedlings depending on the species,
31      but information on physiological traits can be used to predict such differences in some cases.

        January 2005                             9-342       DRAFT-DO NOT QUOTE OR CITE

-------
 1      In some cases, mature trees may be more sensitive to O3 than seedlings due to differences in gas
 2      exchange rates, differences in growth rates, greater cumulative exposure, or the interaction of
 3      ozone with other stresses.
 4
 5      9.9.7  Ecosystem Effects
 6           There is evidence that tropospheric O3 is an important stressor of ecosystems, with
 7      documented impacts on the biotic condition, ecological processes, and chemical/physical nature
 8      of natural ecosystems.  In turn, the effects of O3 on individual plants and processes are scaled up
 9      through the ecosystem affecting processes  such as energy and material flow, inter- and intra-
10      species competition, and  NPP.  Thus, effects on individual keystone species and their associated
11      microflora and fauna, which has been shown experimentally, may cascade through the
12      ecosystem to the landscape level, although it has not yet been demonstrated. By affecting water
13      balance, cold hardiness, tolerance to wind,  and by predisposing plants to insect and disease pests,
14      O3 may even impact the occurrence and impact of natural disturbance (e.g., fire, erosion).
15      Despite the probable occurrence of such effects, there are essentially no instances where
16      ecosystem level, highly integrated studies have conclusively shown that ozone is indeed altering
17      ecosystem structure and/or function.
18           Systematic injury surveys demonstrate that foliar injury occurs on sensitive  species in
19      many regions of the USA. However, the frequent lack of correspondence between foliar
20      symptoms and growth effects means that other methods must be used to estimate the regional
21      effects of O3 on tree growth rates. Investigations  of the radial growth of mature trees in
22      combination with data from many controlled studies with seedlings and a few studies with
23      mature trees suggest that ambient O3 is reducing the growth of mature trees in some locations.
24      Studies using  models based on tree physiology and forest stand dynamics suggest that modest
25      effects of O3 on growth may accumulate over time, and may interact with other stresses.  For
26      mixed-species stands, such models predict  that overall stand growth rate is generally not likely to
27      be affected. However, competitive interactions among species may change as a result of growth
28      reductions of sensitive species.  These results suggest that O3 exposure over decades may be
29      altering the species composition of forests in some regions.
30           Despite increased understanding of possible ecosystem effects of ozone, the data base
31      demonstrating and quantifying the degree to which O3 is altering natural ecosystems is lacking.

        January 2005                              9-343       DRAFT-DO NOT QUOTE OR CITE

-------
 1      Much of the speculation of ozone impact on ecosystems must be inferred from a number of case
 2      studies of forest plot field-based data reporting on a number of different species. One means to
 3      discuss our current knowledge is by listing the areas in which there is lack of knowledge. These
 4      include:
 5           Ecosystem processes. Very little is known about the effects of O3 on water, carbon, and
 6      nutrient cycling, particularly at the stand and community levels. Effects on belowground
 7      ecosystem processes in response to O3, exposure alone and in combination with other stressors
 8      are critical to projections at the watershed and landscape scales. Little is yet known about the
 9      effects of O3 on structural or functional components of soil food webs or how these impacts
10      could affect plant species diversity (Andersen, 2003).
11           Biodiversity and genetic diversity.  The study of genetic aspects of O3 impacts on natural
12      ecosystems has been largely based on correlations and it remains to be shown conclusively
13      whether O3 affects biodiversity or genetic diversity (Pitelka, 1988; Winner et al., 1991; Davison
14      & Barnes, 1998).  Studies of competitive interactions under elevated O3 are needed (Laurence &
15      Andersen, 2003), and re-examination via new sampling of population studies to bring in a time
16      component to previous studies showing spatial variability in population responses to O3 are
17      needed.  These studies could be strengthened by modern molecular methods to quantify impacts
18      on diversity.
19           Natural ecosystem interactions with the atmosphere. Little is known about feedbacks
20      between O3 and climate change on production of volatile organic compounds, which in turn,
21      could affect O3 production (Fuentes et al., 2001).  At moderate to high O3 exposure sites,
22      aberrations in stomatal behavior could significantly affect individual tree water balance of
23      sensitive trees,  and if the sensitive tree species is dominant, hydrologic balance at the watershed
24      and landscape level could be affected. This has not been addressed in any model because O3
25      exposure effects, if included at all in the  modeling effort, have assumed a linear relationship
26      between assimilation and stomatal conductance. Interaction studies with other components of
27      global change (i.e., warming, increasing  atmospheric CO2, N deposition, etc.) or with various
28      biotic stressors are needed to better predict complex interactions likely in  the future (Laurence &
29      Andersen, 2003).  Whether O3 will negate the positive  effects of an elevated CO2 environment
30      on plant carbon and water balances is not yet known nor is it known if these effects will scale-up
        January 2005                              9-344       DRAFT-DO NOT QUOTE OR CITE

-------
 1      through the ecosystem. How O3 affects the progress of pest epidemics and insect outbreaks as
 2      concentrations increase is unclear (Skarby et al., 1998).
 3           Information concerning the impact of O3 on reproductive processes and reproductive
 4      development under realistic field or forest conditions are needed as well as examination of
 5      reproductive effects under interacting pollutants (Black et al., 2000).
 6           Scaling. The vast maj ority of O3 studies of trees have been conducted with young,
 7      immature trees and in trees that have not yet formed a closed canopy. Questions remain as to the
 8      comparability of O3 effects on juvenile and mature trees and on trees grown in the open versus
 9      those in a closed forest canopy in a competitive environment (Chappelka & Samuelson, 1998;
10      Kolb & Matyssek, 2001;  Samuelson & Kelly, 2001). Scaling the effects of O3 across spatial
11      scales is also difficult. Scaling responses of single or a few plants  to effects on communities and
12      ecosystems are complicated matters that will require a combination of manipulative experiments
13      with model ecosystems, community and ecosystem studies along natural O3 gradients, and
14      extensive modeling efforts to project landscape level, regional, national and international
15      impacts of O3. Linking these various studies via impacts on common research quantification
16      across various scales using measures of such factors as leaf area index or spectral reflective data,
17      which could eventually be remotely sensed (Kraft et al., 1996; Panek et al., 2003), would provide
18      powerful new tools for ecologists.
19           Identifying endpoints. In general, methodologies to determine the important values of
20      services and benefits derived from natural ecosystems are lacking. Identifying and quantifying
21      factors that could be used in comprehensive risk assessment for O3 effects on natural ecosystems
22      would increase societal awareness of the importance of protecting  ecosystems (Heck et al.,
23      1998).
24
25      9.9.8  Economics
26           Substantial progress has been made over the past two decades in our understanding of the
27      effects of ozone  and other oxidants on vegetation, particularly for agriculturally-important plant
28      species. The physical and economic effects on agriculture are well documented and provide
29      useful information for the consideration of establishing air quality  standards for crops. Effects
30      on forests and natural ecosystems remain problematic, due to limitations in biological response
        January 2005                              9-345       DRAFT-DO NOT QUOTE OR CITE

-------
 1     data and economic methods. The problem is even more acute for valuing natural ecosystem
 2     service flows.
 3           The current limitations surrounding forests and natural ecosystems present a rich research
 4     agenda.  However, not all research needs are likely to lead to better policies.  Thus, areas of
 5     greatest potential value in terms of regional policy making need to be prioritized. Such
 6     priority-setting can be assisted by sensitivity analyses with existing economic models; by
 7     measuring the changes in economic effects arising from changes in key parameters, it is possible
 8     to identify those research data gaps most likely to affect economic values.
 9
10
       January 2005                              9-346       DRAFT-DO NOT QUOTE OR CITE

-------
  1       REFERENCES

  2       Abeles, F. B.; Biles, C. L.; Dunn, L. J. (1989a) Hormonal regulation and distribution of peroxidase isoenzymes in
  3             the cucurbitaceae. Plant Physiol. 91: 1609-1612.
  4       Abeles, F. B.; Hershberger, W. L.; Dunn, L. J. (1989b) Hormonal regulation, and intracellular localization of a
  5             33-kD cationic peroxidase in excised cucumber cotyledons. Plant Physiol. 89: 664-668.
  6       Adams, R. M. (1986) Agriculture, forestry, and related benefits of air pollution control: a review and some
  7             observations. Am. J. Agric. Econ. 68: 464-472.
  8       Adams, R. M. (1999) On the search for the correct economic assessment method. An editorial comment. Clim.
  9             Change 41: 363-370.
10       Adams, W. C. (2002) Comparison of chamber and face-mask 6.6-hour exposures to ozone on pulmonary function
11             and symptoms responses. Inhalation Toxicol. 14: 745-764.
12       Adams, R. M.; Crocker, T. D. (1989) The agricultural economics of environmental change: some lessons from air
13             pollution. J. Environ. Manage. 28: 295-307.
14       Adams, R. M.; Horst, R. L., Jr. (2003) Future directions in air quality research: economic issues. Environ. Int.
15             29:289-302.
16       Adams, R. M.; Rowe, R. D. (1990) The economic effects of stratospheric ozone depletion on U.S. agriculture:
17             a preliminary assessment. J. Environ. Manage. 30: 321-335.
18       Adams, R. M.; Hamilton, S. A.; McCarl, B. A. (1986) The benefits of pollution control: the case of ozone and U.S.
19             agriculture. Am. J. Agric. Econ. 68: 886-893.
20       Adams, R. M.; Glyer, J. D.; McCarl, B. A. (1988) The NCLAN economic assessment: approach, findings and
21             implications. In: Heck, W. W.; Taylor, O. C.; Tingey, D. T., eds. Assessment of crop loss from air pollutants:
22             proceedings of an international conference; October 1987; Raleigh, NC. New York, NY: Elsevier Applied
23             Science; pp. 473-504.
24       Adams, D. M.; Alig, R. J.; Callaway, J. M.; McCarl, B. A.; Winnett, S. M. (1996) The forest and agricultural sector
25             optimization model (FASOM): model structure and policy application. Portland, OR: U.S. Department of
26             Agriculture, Pacific Northwest Research Station, PNW-RP-495.
27       Adams, R. M.; Kurd, B. H.; Lenhart, S.; Leary, N.  (1998) Effects of global climate change on agriculture: an
28             interpretative review. Climate Research 11:  19-30.
29       Agrawal, M.; Agrawal, S. B. (2000) Effects of air pollution on plant diversity. In: Agrawal, S. B.; Agrawal, M., eds.
30             Environmental Pollution and Plant Responses. Boca Raton, FL: Lewis Publishers; pp.  137-152.
31       Ainsworth, N.; Fumagalli, L; Giorcelli, A.; Mignanego, L.; Schenoe, G.; Vietto, L. (1996) Assessment of EDU stem
32             injections as a technique to investigate the response of trees to ambient ozone in field conditions. Agric.
33             Ecosyst. Environ. 59: 33-42.
34       Ainsworth, E. A.; Davey, P. A.; Bernacchi, C. J.; Dermody, O. C.; Heaton, E. A.; Moore, D. J.; Morgan, P. B.;
35             Naidu,  S. L.; Yoora, H.-S.; Zhu, X.-G.; Curtis, P.  S.; Long, S. P. (2002) A meta-analysis of elevated [CO2]
36             effects on soybean (Glycine max) physiology, growth and yield. Global Change Biol. 8: 695-709.
37       Alscher, R. G.; Wellburn, A. R., eds. (1994) Plant responses  to the gaseous environment. Molecular, metabolic and
3 8             physiological aspects. London, UK: Chapman and Hall.
39       Altshuller, A.  P.; Lefohn, A.  S. (1996) Background ozone in the planetary boundary layer over the United States.
40             J. Air Waste Manage.  Assoc. 46:  134-141.
41       Amiro, B. D.; Gillespie, T. J. (1985) Leaf conductance response ofPhaseolus vulgaris to ozone flux density.
42             Atmos. Environ. 19: 807-810.
43       Amiro, B. D.; Gillespie, T. J.; Thurtell, G. W. (1984) Injury response ofPhaseolus vulgaris to ozone flux density.
44             Atmos. Environ. 18: 1207-1215.
45       Amthor, J. S.; Goulden, M. L.; Munger, J. W.; Wofsy, S. C. (1994) Testing a mechanistic model of forest-canopy
46             mass and energy exchange using eddy correlation: carbon dioxide and ozone uptake by a mixed oak-maple
47             stand. Aust. J. Plant Physiol. 21: 623-651.
48       Anderson, E. (1990) The ethical limitations of the market. Econ. Philos. 6: 179-205.
49       Andersen, C. P. (2000) Ozone stress and changes below-ground: linking root and soil processes. Phyton 40(Special
50             issue 4): 7-12.
51       Andersen, C. P. (2003) Source-sink balance and carbon allocation below ground in plants exposed to ozone.
52             New Phytol. 157: 213-228.
53       Andersen, C. P.; Rygiewicz, P. T. (1995) Allocation of carbon in mycorrhizal Pinusponderosa seedlings exposed to
54             ozone. New Phytol. 131: 471-480.
         January 2005                                  9-347         DRAFT-DO NOT QUOTE OR CITE

-------
  1       Andersen, C. P.; Scagel, C. F. (1997) Nutrient availability alters belowground respiration of ozone-exposed
  2             ponderosa pine. Tree Physiol. 17: 377-387.
  3       Anderson, R. L.; Brown, H. D.; Chevone, B. I.; McCartney, T. C. (1988) Occurence of air pollution symptoms
  4             (needle tip necrosis and chlorotic mottling) on eastern white pine in the southern Appalachian Mountains.
  5             Plant Dis. 72: 130-132.
  6       Andersen, C. P.; Hogsett, W. E.; Wessling, R.; Plocher, M. (1991) Ozone decreases spring root growth and root
  7             carbohydrate content in ponderosa pine the year following exposure. Can. J. For. Res. 21: 1288-1291.
  8       Andersen, C. P.; Wilson, R.; Plocher, M.; Hogsett, W. E. (1997) Carry-over effects of ozone on root growth and
  9             carbohydrate concentrations of ponderosa pine seedlings. Tree Physiol. 17: 805-811.
10       Andersen, C. P.; Hogsett, W. E.; Plocher, M.; Rodecap, K.; Lee, E. H. (2001) Blue wild-rye grass competition
11             increases the effect of ozone on ponderosa pine seedlings. Tree Physiol. 21:319-327.
12       Antonielli, M.; Pasqualini,  S.; Ederli, L.; Batini, P.; Moscatello, S.; Loreto, F. (1997) Physiological characteristics of
13             tobacco cultivars with contrasting sensitivity to ozone. Environ. Exp. Bot. 38:  271-277.
14       Anttonen, S.; Sutinen, M. L.; Heagle, A. S. (1996) Ultrastructure and some plasma membrane characteristics of
15             ozone-exposed loblolly pine needles. Physiol. Plant. 98: 309-319.
16       Aono, M.; Saji, H.; Fujiyama, K.;  Sugita, M.; Kondo, N.; Tanaka, K. (1995) Decrease in activity of glutathione
17             reductase enhances paraquat sensitivity in transgenic Nicotiana tabacum. Plant Physiol. 107: 645-648.
18       Arbaugh, M. J.; Miller, P. R.; Carroll, J. J.; Takemoto, B. L.; Proctor, T. (1998) Relationships of ozone exposure to
19             pine injury in the Sierra Nevada and San Bernardino Mountains of California, USA. Environ. Pollut.
20             101:291-301.
21       Arbaugh, M.; Bytnerowicz, A.; Grulke, N.; Fenn, M.; Poth, M.; Temple, P.; Miller, P. (2003) Photochemical smog
22             effects in mixed conifer forests along a natural gradient of ozone and nitrogen deposition in the
23             San Bernardino Mountains. Environ. Int. 29: 401-406.
24       Arie'ns, E. J.; Simonis, A. M.; Offermeier, J. (1976) Introduction to general toxicology. New York, NY:
25             Academic Press.
26       Arndt, U. (1995) Air pollutants and pheromones - a problem? Chemosphere 30: 1023-1031.
27       Arp, W. J.; Drake, B.  G. (1991) Increased photosynthetic capacity ofScirpus olneyi after 4 years of exposure to
28             elevated CO2. Plant Cell Environ. 14: 1003-1006.
29       Asada, K. (1999) The water-water cycle in chloroplasts: scavenging of active oxygens and dissipation of excess
30             photons. Ann. Rev. Plant Physiol. Mol. Biol. 50: 601-639.
31       Ashenden, T. W.; Bell, S. A.; Rafarel, C. R. (1995) Responses of white clover to gaseous pollutants and acid mist:
32             implications for setting critical levels and loads. New Phytol. 130: 89-96.
33       Ashenden, T. W.; Bell, S. A.; Rafarel, C. R. (1996) Interactive effects of gaseous air pollutants and acid mist on two
34             major pasture grasses. Agric. Ecosyst. Environ. 57: 1-8.
3 5       Ashmore, M. R. (1984) Effects of ozone on vegetation in the United Kingdom. In: Grennfelt, P., ed. Proceedings of
3 6             the international workshop  on the evaluation of the effects of photochemical oxidants on human health,
37             agricultural crops, forestry, materials and visibility; February-March; Goteborg, Sweden. Goteborg, Sweden:
3 8             Swedish Environmental Research Institute; pp. 92-104; report no. IVL-EM 1570.
39       Ashmore, M. R.; Ainsworth, N. (1995) The effects of ozone and cutting on the species composition of artificial
40             grassland communities.  Funct. Ecol. 9: 708-712.
41       Ashmore, M.; Fuhrer, J. (2000) New directions: use and abuse of the AOT40 concept. Atmos. Environ.
42             34: 1157-1158.
43       Ashmore, M. R.; Bell, J. N. B.; Mimmack, A. (1988) Crop growth along a gradient of ambient air pollution.
44             Environ. Pollut. 53:  99-121.
45       Ashmore, M. R.; Thwaites, R. H.; Ainsworth, N.; Cousins, D. A.; Power,  S. A.; Morton, A. J. (1995) Effects of
46             ozone on calcareous grassland communities. Water Air Soil Pollut. 85: 1527-1532.
47       Atkinson, R. (1990) Gas-phase tropospheric chemistry of organic compounds: a review. Atmos. Environ. Part A
48             24: 1-41.
49       Auh, C.-K.; Murphy,  T. M. (1995) Plasma membrane redox enzyme is involved in the synthesis of O2" and H2O2 by
50             Phytophthora e//c/tor-stimulated rose cells. Plant Physiol. 107: 1241-1247.
51       Awmack, C. S.; Harrington, R.; Lindroth, R. L. (2004) Aphid individual performance may not predict population
52             responses to elevated CO2 or O3. Global Change Biol. 10: 1414-1423.
53       Ayres, P. G. (1991) Growth responses induced by pathogens and other stresses. In: Mooney, H. A.; Winner, W. E.;
54             Pell, E. J., eds. Responses of plants to multiple stresses. San Diego, CA: Academic Press; pp. 227-248.
         January 2005                                   9-348         DRAFT-DO NOT QUOTE OR CITE

-------
  1       Azevedo, R. A.; Alas, R. M; Smith, R. I; Lea, P. J. (1998) Response of antioxidant enzymes to transfer from
  2             elevated carbon dioxide to air and ozone fumigation, in the leaves and roots of wild-type and a
  3             catalase-deficient mutant of barley. Physiol. Plant. 104: 280-292.
  4       Back, J.; Vanderklein, D. W.; Topa, M. A. (1999) Effects of elevated ozone on CO2 uptake and leaf structure in sugar
  5             maple under two light environments. Plant Cell Environ. 22: 137-147.
  6       Bahl, A.; Kahl, G. (1995) Air pollutant stress changes the steady-state transcript levels of three photosynthesis genes.
  7             Environ. Pollut. 88: 57-65.
  8       Bahl, A.; Loitsch, S. M.; Kahl, G. (1995) Transcriptional activation of plant defence genes by short-term air pollutant
  9             stress. Environ. Pollut. 89: 221-227.
10       Baker, T. R.; Allen, H. L.; Schoeneberger, M. M.; Kress, L. W. (1994)  Nutritional response of loblolly pine exposed
11             to ozone and simulated acid rain. Can. J. For. Res. 24: 453-461.
12       Baker, C. J.; Harmon, G. L.; Glazener, J. A.; Orlandi, E. W. (1995) A noninvasive technique for monitoring
13             peroxidative and H2O2-scavenging activities during interactions between bacterial plant pathogens and
14             suspension cells. Plant Physiol. 108: 353-359.
15       Balaguer, L.; Barnes, J. D.; Panicucci, A.; Borland, A. M. (1995) Production and utilization of assimilates in wheat
16             (Triticum aestivum L.) leaves exposed to elevated O3 and/or CO2. New Phytol. 129: 557-568.
17       Ball, J. T. (1987) Calculations related to gas exchange. In: Zeiger, E.; Farquhar, G. D.; Cowan, I. R. Stomatal
18             Function. Stanford, CA: Stanford Press; pp. 445-476, pp.  G-47.
19       Ball, G. R.; Benton, J.; Palmer-Brown, D.; Fuhrer, J.; Skarby, L.; Gimeno, B. S.; Mills, G. (1998) Identifying factors
20             which modify the effects of ambient ozone on white clover (Trifolium repens) in Europe. Environ. Pollut.
21             103:7-16.
22       Ball, G. R.; Palmer-Brown, D.; Fuhrer, J.; Skarby, L.;  Gimeno, B. S.; Mills, G. (2000) Identification of non-linear
23             influences on the seasonal ozone dose-response of sensitive and  resistant clover clones using artificial neural
24             networks. Ecol. Model. 129: 153-168.
25       Balls, G. R.; Palmer-Brown, D.; Cobb, A. H.; Sanders, G. E.  (1995) Towards unrevelling the complex interactions
26             between microclimate, ozone dose, and ozone injury in clover. Water Air Soil. Pollut. 85:  1467-1472.
27       Balls, G. R.; Palmer-Brown, D.; Sanders, G. E. (1996) Investigating microclimatic influences on ozone injury in
28             clover (Trifolium subterraneum) using artificial neural networks. New Phytol. 132: 271-280.
29       Banerjee, S. K.; Mudd, J. B. (1992) Reaction of ozone with glycophorin in solution and in lipid vesicles.
30             Arch. Biochem. Biophys. 295: 84-89.
31       Barbo, D. N.; Chappelka, A. H.; Somers, G. L.; Miller-Goodman, M. S.; Stolte, K. (1998) Diversity of an early
32             successional plant community as influenced by ozone. New Phytol. 138: 653-662.
33       Barbo, D. N.; Chappelka, A. H.; Somers, G. L.; Miller-Goodman, M. S.; Stolte, K. (2002) Ozone impacts on loblolly
34             pine (Firms taeda L.) grown in a competitive environment. Environ. Pollut.  116: 27-36.
35       Barnes,  J. D.; Pfirrmann, T. (1992) The influence of CO2 and O3, singly and in combination, on gas exchange, growth
36             and nutrient status of radish (Raphanus sativus L.). New Phytol.  121: 403-412.
37       Barnes,  J. D.; Wellburn, A. R. (1998) Air pollutant combinations. In: De Kok, L. J.; Stolen, L, eds. Responses of
3 8             plant metabolism to air pollution and global change. Leiden, The Netherlands: Backhuys Publishers;
39             pp.  147-164.
40       Barnes,  J. D.; Ollerenshaw, J. H.; Whitfield, C. P. (1995) Effects of elevated CO2 and/or O3 on growth, development
41             and physiology of wheat (Triticum aestivum L.). Global Change  Biol. 1: 129-142.
42       Barnes,  J.; Zheng, Y.; Lyons, T. (2002) Plant resistance to ozone: the role of ascorbate. In: Omasa, K.; Saji, H.;
43             Youssefian, S.; Kondo, N., eds. Air pollution and plant biotechnology - prospects for phytomonitoring and
44             phytoremediation. Springer-Verlag; pp. 235-252.
45       Barrett,  S. C. H.;  Bush, E. J. (1991) Population processes in plants and  the evolution of resistance to gaseous air
46             pollutants. In: Taylor, G. E., Jr.; Pitelka, L. F.; Clegg, M.  T., eds. Ecological genetics and air pollution.
47             New York, NY: Springer-Verlag; pp. 137-165.
48       Bartholomay, G. A.; Eckert, R. T.; Smith, K. T. (1997) Reductions in tree-ring widths of white pine following ozone
49             exposure at Acadia National Park, Maine, USA. Can. J. For. Res. 27: 361-368.
50       Batts, G. R.; Morison, J. I. L.; Ellis, R. H.; Hadley, P.; Wheeler,  T. R. (1997) Effects of CO2 and temperature on
51             growth and yield of crops of winter wheat over four seasons. Eur. J. Agron.  7: 43-52.
52       Bauer, M. R.; Hultman, N. E.; Panek, J. A.; Goldstein, A. H.  (2000) Ozone deposition to a ponderosa pine plantation
53             in the Sierra Nevada Mountains (CA): a comparison of two different climatic years. J. Geophys. Res. [Atmos.]
54             105:22,123-22,136.
         January 2005                                   9-349         DRAFT-DO NOT QUOTE OR CITE

-------
  1       Baumbusch, L. O.; Eiblmeier, M; Schnitzler, J.-P.; Heller, W.; Sandermann, H., Jr.; Polle, A. (1998) Interactive
  2             effects of ozone and low UV-B radiation on antioxidants in spruce (Picea abies) and pine (Pinus sylvestris).
  3             Physiol. Plant. 104: 248-254.
  4       Baumgardner, R. E.; Edgerton, E. S. (1998) Rural ozone across the eastern United States: analysis of CASTNet data,
  5             1988-1995. J. Air Waste Manage. Assoc. 48: 674-688.
  6       Baumgarten, M.; Werner, H.; Haberle, K.-H.; Emberson, L. D.; Fabian, P.; Matyssek, R. (2000) Seasonal ozone
  7             response of mature beech trees (Fagus sylvatica) at high altitude in the Bavarian Forest (Germany) in
  8             comparison with young beech grown in the field and in phytotrons. Environ. Pollut. 109: 431-442.
  9       Beare, J. A.; Archer, S. A.; Bell, J. N. B. (1999a) Effects ofMelampsora leaf rust disease and chronic ozone exposure
10             on poplar. Environ. Pollut.  105: 419-426.
11       Beare, J. A.; Archer, S. A.; Bell, J. N. B. (1999b) Marssonina leafspot disease of poplar under elevated ozone:
12             pre-fumigated host and in vitro studies. Environ. Pollut. 105: 409-417.
13       Beligni, M. V.; Lamattina, L. (2001) Nitric oxide in plants: the history is just beginning. Plant Cell Environ.
14             24:267-278.
15       Beligni, M.; Lamattina, L. (2002) Nitric oxide interferes with plant photo-oxidative stress by detoxifying reactive
16             oxygen species. Plant Cell Environ. 25: 737-748.
17       Bell, J. N. B.; Treshow, M., eds. (2002) Air pollution and plant life. 2nd ed. Chichester, United Kingdom: John Wiley
18             & Sons Ltd.
19       Bell, J. N. B.; Ashmore, M. R.;  Wilson, G. B. (1991) Ecological genetics and chemical modifications of the
20             atmosphere. In: Taylor, G. E.,  Jr.; Pitelka, L. F.; Clegg, M. T., eds. Ecological genetics and air pollution.
21             New York, NY: Springer-Verlag; pp. 33-59.
22       Bender, J.; Weigel, H. J.; Wegner, U.; Jager, H. J. (1994) Response of cellular antioxidants to ozone in wheat flag
23             leaves at different stages of plant development.  Environ. Pollut. 84: 15-21.
24       Bender, J.; Herstein, U.; Black,  C. R. (1999) Growth and yield responses of spring wheat to increasing carbon
25             dioxide, ozone and physiological stresses: a statistical analysis of 'ESPACE-wheat'  results. Eur.  J. Agron.
26             10: 185-195.
27       Benes, S. E.; Murphy, T. M.; Anderson, P. D.; Houpis, J. L. J. (1995) Relationship of antioxidant enzymes  to ozone
28             tolerance in branches of mature ponderosa pine (Pinusponderosd) trees exposed to long term, low
29             concentration, ozone fumigation and acid precipitation. Physiol. Plant. 94: 124-134.
30       Bennett, J. H. (1979) Foliar exchange of gases. In:  Heck, W. W.; Krupa, S. V.; Linzon, S.  N., eds. Methodology for
31             the assessment of air pollution effects on vegetation:  a handbook from a specialty conference; April  1978;
32             Minneapolis, MN. Pittsburgh, PA: Air Pollution Control Association; pp. 10-1 -10-29.
33       Bennett, J. P.; Anderson, R. L.;  Mielke, M. L.; Ebersole, J. J. (1994) Foliar injury air pollution surveys of eastern
34             white-pine (Pinus-strobus L.) - a review. Environ. Monit. Assess. 30:  247-274.
35       Benoit, L. F.; Skelly, J. M.; Moore, L. D.; Dochinger, L. S. (1982) Radial growth reductions of Pinus strobus L.
3 6             correlated with foliar ozone sensitivity as an indicator of ozone-induced losses in eastern forests. Can. J. For.
37             Res. 12: 673-678.
38       Benoit, L. F.; Skelly, J. M.; Moore, L. D.; Dochinger, L. S. (1983) The influence of ozone onPinus strobus L. pollen
39             germination. Can. J. For. Res.  13: 184-187.
40       Benton, J.; Fuhrer, J.; Gimeno, B. S.; Skarby, L.; Sanders, G. (1995) Results from the UN/ECE ICP-crops indicate the
41             extent of exceedance of the critical levels of ozone in Europe. Water Air Soil Pollut. 85: 1473-1478.
42       Benton, J.; Fuhrer, J.; Gimeno, B. S.; Skarby, L.; Palmer-Brown, D.; Ball, G.; Roadknight, C.; Sanders-Mills, G.
43             (1996) The critical level of ozone for visible injury on crops and natural vegetation (ICP-Crops). In:
44             Karenlampi, L.; Skarby, L., eds. Critical levels for ozone in Europe: testing and finalizing the concepts:
45             UN-ECE workshop report. UN-ECE Convention on Long-Range Transboundary Air Pollution; April; Kuopio,
46             Finland. Kuopio, Finland: University of Kuopio, Department of Ecology and Environmental Science;
47             pp. 44-57.
48       Benton, J.; Fuhrer, J.; Gimeno, B. S.; Skarby, L.; Palmer-Brown, D.; Ball, G.; Roadknight, C.; Mills, G. (2000)
49             An international cooperative programme indicates the widespread occurrence of ozone injury on crops.
50             Agric. Ecosyst. Environ. 78: 19-30.
51       Berg, B.; Staaf, H. (1980) Decomposition rate and  chemical changes of Scots pine needle litter.  II. Influence of
52             chemical composition. Ecol. Bull. 32: 373-390.
53       Bergmann, E.; Bender, J.; Weigel, H.-J. (1999) Ozone threshold doses and exposure-response relationships for the
54             development of ozone injury symptoms in wild plant species. New Phytol. 144: 423-435.
55       Bergstrom, J. C. (1990) Concepts and measures of the economic value of environmental quality: a review. J. Environ.
56             Manage. 31:215-228.
         January 2005                                    9-350        DRAFT-DO NOT QUOTE OR CITE

-------
  1       Bergweiler, C. I; Manning, W. J. (1999) Inhibition of flowering and reproductive success in spreading dogbane
  2             (Apocynum androsaemifolium) by exposure to ambient ozone. Environ. Pollut. 105: 333-339.
  3       Bermejo, V.; Gimeno, B. S.; Granados, I.; Santamaria, J.; Irigoyen, J. J.; Bermejo, R.; Porcuna, J. L.; Mills, G. (2002)
  4             Investigating indices to explain the impacts of ozone on the biomass of white clover (Trifolium repens L. cv.
  5             Regal) at inland and coastal sites in Spain. New Phytol. 156: 43-55.
  6       Berrang, P.; Karnosky, D. F.; Mickler, R. A.; Bennett, J. P. (1986) Natural selection for ozone tolerance inPopulus
  1             tremuloides. Can. J. For. Res. 16: 1214-1216.
  8       Berrang, P.; Karnosky, D. F.; Bennett, J. P. (1989) Natural selection for ozone tolerance in Populus tremuloides: field
  9             verification. Can. J. For. Res. 19: 519-522.
10       Bertrand, A.; Robitaille, G.; Nadeau, P.; Castonguay, Y. (1999)  Influence of ozone on cold acclimation in sugar
11             maple seedlings. Tree Physiol. 19: 527-534.
12       Beyers, J. L.; Riechers, G. H.; Temple, P.  J. (1992) Effects of long-term ozone exposure and drought on the
13             photosynthetic capacity of ponderosa pine (Pinusponderosa Laws.). New Phytol. 122: 81-90.
14       Bichele, L; Moldau, H.; Padu, E. (2000) Estimation of plasmalemma conductivity to ascorbic acid in intact leaves
15             exposed to ozone. Physiol. Plant. 108: 405-412.
16       Bielenberg, D. G.; Lynch, J. P.; Pell, E. J. (2001) A decline in nitrogen availability affects plant responses to ozone.
17             New Phytol. 151:413-425.
18       Binkley, D. (1992) Sensitivity of forest soils in the western U.S. to acidic deposition. In: Olson, R. K.; Binkley, D.;
19             Bohm, M., eds. The Response of Western Forests to Air Pollution. New York, NY: Springer-Verlag;
20             pp. 153-182.
21       Birecka, H.; Catalfamo, J.; Urban, P. (1976) Cell isoperoxidases in sweet potato plants in relation to mechanical
22             injury and ethylene. Plant Physiol. 57: 74-79.
23       Black, V. J.; Black, C.  R.; Roberts,  J. A.; Stewart, C. A. (2000) Impact of ozone on the reproductive development of
24             plants. New Phytol. 147: 421-447.
25       Blankenship, E. E.; Stefanski, L. A. (2001) Statistical estimation of ozone exposure metrics.  Atmos.  Environ.
26             35:4499-4510.
27       Boerner, R. E. J.; Rebbeck, J. (1995) Decomposition and nitrogen release from leaves of three hardwood species
28             grown under elevated O3 and/or CO2. Plant Soil 170: 149-157.
29       Bohm, M.; McCune, B.; Vandetta, T. (1995) Ozone regimes in or near forests of the western United States: I. regional
30             patterns. J. Air Waste Manage. Assoc. 45: 235-246.
31       Bonello, P.; Heller, W.; Sandermann, H., Jr. (1993) Ozone effects on root-disease susceptibility and defence
32             responses in mycorrhizal and non-mycorrhizal seedlings of Scots pine  (Pinus sylvestris L.). New Phytol.
33             124:653-663.
34       Booker, F. L.  (2000) Influence of carbon dioxide enrichment, ozone and nitrogen fertilization on cotton
35             (Gossypium hirsutum L.) leaf and root composition. Plant Cell Environ. 23: 573-583.
36       Booker, F. L.; Reid, C. D.; Brunschon-Harti, S.; Fiscus, E. L.; Miller, J. E. (1997) Photosynthesis and
37             photorespiration in soybean [Glycine max (L.) Merr] chronically exposed to elevated carbon  dioxide and
38             ozone. J. Exp. Bot.  48: 1843-1852.
39       Border, K.;  Ceulemans, R.; De Temmerman, L. (2000) Effects of tropospheric ozone on woody plants. In: Agrawal,
40             S. B.; Agrawal, M., eds. Environmental pollution and plant responses;  pp. 153-182. Boca Raton, FL:
41             Lewis Publishers.
42       Border, K.;  Ceulemans, R.; De Temmerman, L. (2000a) Effects  of ozone exposure on growth and photosynthesis of
43             beech seedlings  (Fagus sylvaticd). New Phytol. 146: 271-280.
44       Border, K.; DeTemmerman, L.;  Ceulemans, R. (2000b) Effects of ozone exposure in open-top chambers on poplar
45             (Populus nigra)  and beech (Fagus sylvaticd): a comparison. Environ. Pollut. 109: 509-516.
46       Border, K.; Vandermeiren, K.; De Temmerman, L.; Ceulemans, R. (200la) Growth, photosynthesis  and ozone uptake
47             of yound beech (Fagus sylvatica L.) in response to different ozone exposures. Trees-Struct. Funct. 15: 75-82
48       Border, K.; Dekelver, G.; De Temmerman, L.; Ceulemans, R. (200 Ib) Stem injection of Populus nigra with EDU to
49             study ozone effects under field conditions. Environ. Pollut. Ill: 199-208.
50       Bosac, C.; Black, V.  J.; Roberts, J. A.; Black, C. R. (1998) The impact of ozone on seed yield and quality and
51             seedling vigour in oilseed rape (Brassica napus L.). J. Plant Physiol.  153: 127-134.
52       Bosley, A.; Peterson, R.; Rebbeck, J. (1998) The resistance of the moss Polytrichum commune to acute exposure of
53             simulated acid rain or ozone  compared to two fern species: spore germination. Bryologist 101: 512-518.
54       Bowler, C.; Van Montagu, M.; Inze, D. (1992) Superoxide dismutase and stress tolerance. Ann. Rev. Plant Physiol.
55             Mol.Biol. 43:83-116.
         January 2005                                   9-3 51        DRAFT-DO NOT QUOTE OR CITE

-------
  1       Bradshaw, A. D.; McNeilly, T. (1991) Evolution in relation to environmental stress. In: Taylor, G. E., Jr.; Pitelka,
  2             L. F.; Clegg, M. T., eds. Ecological genetics and air pollution. New York, NY: Springer-Verlag; pp.  11-31.
  3       Braun, S.; Fltickiger, W. (1995) Effects of ambient ozone on seedlings ofFagus sylvatica L. and Picea abies (L.)
  4             Karst. New Phytol. 129: 33-44.
  5       Braun, S.; Rihm, B.; Schindler, C.; Fluckiger, W. (1999) Growth of mature beech in relation to ozone and nitrogen
  6             deposition:  an epidemiological approach. Water Air Soil Pollut. 116: 357-364.
  7       Brennan, E. G.; Leone, I. A.; Clarke, B. B. (1987) EDU: a chemical for evaluating ozone foliar injury and yield
  8             reduction in field-grown crops. Int. J. Trop. Plant Dis. 5: 35-42.
  9       Bringmark, E.; Bringmark, L. (1995) Disappearance of spatial variability and structure in forest floors - a distinct
10             effect of air pollution? Water Air SoilPollut. 85: 761-766.
11       Broadmeadow, M. (1998) Ozone and forest trees. New Phytol. 139: 123-125.
12       Broadmeadow, M. S. J.; Heath, J.; Randle, T. J. (1999) Environmental limitations to O3 uptake - some key results
13             from young trees growing at elevated CO2 concentrations. Water Air Soil Pollut. 116: 299-310.
14       Broadmeadow, M. S. J.; Jackson, S. B. (2000) Growth responses of Quercuspetraea, Fraxinus excelsior and Pinus
15             sylvestris to elevated carbon dioxide, ozone  and water supply.  New Phytol.  146: 437-451.
16       Brook, J. R.; DiGiovanni, F.; Cakmak, S.; Meyers, T. P. (1997) Estimation of dry deposition velocity using inferential
17             models and site-specific meteorology-uncertainty due to siting of meteorological towers. Atmos. Environ.
18             31:3911-3919.
19       Brown, V. C. (1995) Insect herbivores and gaseous air pollutants - current knowledge and predictions. In:
20             Harrington, R.; Stork, N. E., eds. Insects in a changing environment:  17th symposium of the Royal
21             Entomological Society. New York, NY: Academic Press; pp. 219-249.
22       Brown, K. A.; Roberts, T. M. (1988) Effects of ozone  on foliar leaching in Norway spruce (Picea abies L. Karst):
23             confounding factors due to NOX production during ozone  generation. Environ. Pollut. 55: 55-73.
24       Brown, V. C.; Ashmore, M. R.; McNeill, S. (1993) Experimental investigations of the effects of air pollutants on
25             aphids on coniferous trees. Forstwiss. Centralbl. 112: 129-132.
26       Brunschon-Harti, S.; Fangmeier, A.; Jager, H.-J. (1995) Influence of ozone and ethylenediurea (EDU) on growth and
27             yield of bean (Phaseolus vulgaris L.) in open-top field chambers. Environ. Pollut. 90: 89-94.
28       Buchanan, B. B.; Gruissem, W.; Jones, R. L., eds. (2000) Biochemistry and molecular biology of plants. Rockville,
29             MD: American Society of Plant Physiologists.
30       Bungener, P.; Balls, G. R.; Nussbaum, S.; Geissmann, M.; Grub, A.;  Fuhrer, J.  (1999a) Leaf injury characteristics of
31             grassland species exposed to ozone in relation to soil moisture condition and vapour pressure deficit.
32             New Phytol. 142: 271-282.
33       Bungener, P.; Nussbaum, S.; Grub, A.; Fuhrer, J. (1999b) Growth response of grassland species to ozone in relation
34             to soil moisture and plant strategy. New Phytol. 142: 283-293.
3 5       Burkey, K. O. (1999) Effects of ozone on apoplast/cytoplasm partitioning of ascorbic acid in snap bean. Physiol.
36             Plant. 107:  188-193.
37       Bussotti, F.; Ferretti, M. (1998) Air pollution, forest condition and forest decline in Southern Europe. Environ. Pollut.
38              101:49-65.
3 9       Bytnerowicz, A. (2002) Physiological/ecological interactions between ozone and nitrogen deposition in forest
40             ecosystems. Phyton 42: 13-28.
41       Bytnerowicz, A.; Turunen, M. (1994) Effects of ozone exposures on epicuticular wax of ponderosa pine needles. In:
42             Percy, K. E.; Cape, J. N.; Jagels, R.; Simpson, C. J., eds. Air Pollutants and the Leaf Cuticle. New York, NY:
43             Springer-Verlag; pp. 305-313. (NATO ASI Series; v. G 36).
44       Bytnerowicz, A.; Olszyk, D. M.; Fox, C. A.; Dawson,  P. J.; Kats, G.; Morrison, C. L.; Wolf, J. (1988) Responses of
45             desert annual plants to ozone and water stress in an in situ experiment. JAPCA 38: 1145-1151.
46       Bytnerowicz, A.; Olszyk, D. M.; Huttunen, S.; Takemoto, B. K.  (1989) Effects  of photochemical smog on growth and
47             injury and gas exchange of pine seedlings. Can. J. Bot.  67: 2175-2181.
48       Bytnerowicz, A.; Poth, M.; Takemoto, B. K. (1990) Effects of photochemical smog and mineral nutrition on
49             ponderosa pine seedlings. Environ. Pollut. 67: 233-248.
50       Bytnerowicz, A.; Fenn, M. E.; Miller, P. R.; Arbaugh,  M. J. (1999) Wet and dry pollutant deposition to the  mixed
51             conifer forest. In: Miller, P. R.; McBride, J. R.,  eds. Oxidant Air Pollution Impacts in the Montane Forests of
52             Southern California: a Case Study of the San Bernardino Mountains.  New York, NY: Springer; pp. 235-269.
53             (Ecological Series, no. 134).
54
         January 2005                                   9-3 52         DRAFT-DO NOT QUOTE OR CITE

-------
  1       Bytnerowicz, A.; Godzik, B.; Fr^czek, W.; Grodziriska, K.; Krywult, M; Badea, O.; Barancok, P.; Blum, O.;
  2             Cerny, M.; Godzik, S.; Mankovska, B.; Manning, W.; Moravcik, P.; Musselman, R.; Oszlanyi, I;
  3             Postelnicu, D. Szdzuj, J.; Varsavova, M.; Zota, M. (2002) Distribution of ozone and other air pollutants in
  4             forests of the Carpathian Mountains in central Europe. Environ. Pollut. 116: 3-25.
  5       Bytnerowicz, A.; Tausz, M.; Alonso, R.; Jones, D.; Johnson, R.; Grulke, N. (2002) Summer-time distribution of air
  6             pollutants in Sequoia National Park, California. Environ. Pollut. 118: 187-203.
  7       Cairnes, M. S.; Lackey, R. T. (1992) Biodiversity and management of natural resources: the issues. Fisheries
  8             17:6-10.
  9       Callaway, J. M.; Darwin, R. F.; Nesse, R.  J. (1986) Economic valuation of acidic deposition damages: preliminary
10             results from the 1985 NAPAP assessment. In: Proceedings of International Symposium on Acidic
11             Precipitation; September; Muskoka, Ontario, Canada. Richland, WA: U. S. Environmental Protection Agency;
12             Pacific Northwest Laboratory. Water Air Soil Pollut.  31: 1019-1034.
13       Campbell, S.; Temple, P.; Pronos, J.; Rochefort, R.; Andersen, C. (2000) Monitoring for ozone injury in west coast
14             (Oregon, Washington, California) forests in  1998. Portland, OR: U.S. Department of Agriculture, Forest
15             Service, Pacific Northwest Research Station; general technical report no. PNW-GTR-495. Available:
16             http://www.fs.fed.us/pnw/gtrs.htm [11  April, 2003].
17       Cao, W.; Tibbitts, T. W.; Wheeler, R. M. (1994) Carbon dioxide interactions with irradiance and temperature in
18             potatoes. In: MacElroy, R. D. Life sciences and space research XXV(3): natural and artificial ecosystems:
19             proceedings of the topical meeting of the COSPAR Interdisciplinary Scientific Commission F (Meeting F4)
20             of the COSPAR twenty-ninth plenary meeting; August-September, 1992; Washington, DC. New York, NY:
21             Committee on Space Research; pp.  243-250. (Advances in space research: v. 14, no. 11).
22       Cape, J. N. (1996) Surface wetness and pollutant deposition. In: Kerstiens, G., ed. Plant cuticles: an integrated
23             functional approach. Oxford, United Kingdom: BIOS Scientific Publishers; pp. 283-300. (Environmental plant
24             biology series).
25       Cardoso-Vilhena, J.; Barnes, J. (2001) Does nitrogen supply affect the response of wheat (Triticum aestivum cv.
26             Hanno) to the combination of elevated CO2 and O3? J. Exp. Bot. 52: 1901-1911.
27       Cardoso-Vilhena, J.; Balaguer, L.; Daymond, J.; Barnes, J. D. (1998) Amelioration of ozone damage by elevated
28             CO2. In: DeKok, L. J.; Stolen, I., eds. Responses of plant metabolism to air pollution and global change.
29             Leiden, The Netherlands: Backhuys Publishers; pp. 281-282.
30       Carey, W. A.; Kelley, W. D. (1994)  Interaction of ozone exposure and Fusarium subglutinans inocculation on growth
31             and disease development of loblolly pine seedlings. Environ. Pollut. 84: 35-43.
32       Carey, W. A.; Kelley, W. D. (1994)  Effects of ozone exposure and Fusarium subglutinans inoculation on growth and
33             disease development of loblolly pine seedlings. Environ. Pollut. 84: 35-43.
34       Carnahan, J. E.; Jenner, E. L.; Wat, E. K. W. (1978) Prevention of ozone injury to plants by a new protectant
35             chemical. Phytopathology 68: 1225-1229.
36       Carpenter, R. A.; Dixon,  J. A. (1985) Ecology meets economics: a guide to sustainable development. Environment
37             27:6-32.
38       Carroll, J. J.; Miller, P. R.; Pronos, J. (2003) Historical perspectives on ambient ozone and its effects on the Sierra
39             Nevada. In: Bytnerowicz, A.;  Arbaugh, M. J.; Alonso, R., eds. Ozone air pollution in the Sierra Nevada:
40             distribution and effects on forests; pp. 33-54. New York, NY: Elsevier. (Developments in Environmental
41             Science: v. 2).
42       Castillo, F. J.; Greppin, H. (1986) Balance between anionic and cationic extracellular peroxidase activities in Sedum
43             album leaves after ozone exposure. Analysis by high-performance liquid chromatography. Physiol. Plant.
44             68:201-208.
45       Castillo, F. J.; Heath, R. L. (1990) CA2+ transport in membrane vesicles from pinto bean leaves and its alteration after
46             ozone exposure. Plant Physiol. 94: 788-795.
47       Castillo, F. J.; Penel, C.; Greppin, H. (1984) Peroxidase release induced by ozone in Sedum album leaves:
48             involvement of Ca2+. Plant Physiol. 74: 846-851.
49       Castillo, F. J.; Miller,  P. R.; Greppin, H. (1987) Extracellular biochemical markers of photochemical oxidant air
50             pollution damage to Norway spruce. Experientia (Basel) 43: 111-115.
51       Castle, E. N. (1993) A pluralistic, pragmatic and evolutionary approach to natural resource management. For. Ecol.
52             Manage. 56: 279-295.
53       Cazale, A.-C.; Rouet-Mayer, M.-A.;  Barbier-Brygoo, H.; Mathieu, Y.; Lauriere, C. (1998) Oxidative burst and
54             hypoosmotic stress in tobacco cell suspensions.  Plant Physiol. 116: 659-669.
55       Chappelka, A. H. (2002)  Reproductive development of blackberry (Rubus cuneifolius) as influenced by ozone.
56             New Phytol. 155: 249-255.
         January 2005                                   9-353        DRAFT-DO NOT QUOTE OR CITE

-------
  1       Chappelka, A. H., Ill; Chevone, B. I. (1986) White ash seedling growth response to ozone and simulated acid rain.
  2             Can. J. For. Res. 16: 786-790.
  3       Chappelka, A. H.; Freer-Smith, P. H. (1995) Predisposition of trees by air pollutants to low temperatures and moisture
  4             stress. Environ. Pollut. 87:  105-117.
  5       Chappelka, A. H., Ill; Chevone, B. I.; Burk, T. E. (1985) Growth response of yellow-poplar (Liriodendron tulipifera
  6             L.) seedlings to ozone, sulfur dioxide, and simulated acidic precipitation, alone and in combination. Environ.
  7             Exp. Bot. 25: 233-244.
  8       Chappelka, A.; Renfro, J.; Somers, G.; Nash, B.  (1997) Evaluation of ozone injury on foliage of black cherry
  9             (Prunus serotina) and tall milkweed (Asdepias exaltata) in Great Smoky Mountains National Park. Environ.
10             Pollut. 95: 13-18.
11       Chappelka, A. H.; Samuelson, L. J. (1998) Ambient ozone effects on forest trees of the eastern United States:
12             a review. New Phytol. 139: 91-108.
13       Chappelka, A.; Skelly, J.; Somers, G.; Renfro, J.; Hildebrand, E. (1999a) Mature black cherry used as a bioindicator
14             of ozone injury. Water Air  Soil Pollut. 116: 261-266.
15       Chappelka, A.; Somers, G.; Renfro, J. (1999b) Visible ozone injury on forest trees in Great Smoky Mountains
16             National Park, USA. Water Air Soil Pollut. 116: 255-260.
17       Chappelka, A. H.; Neufeld, H. S.; Davison, A. W.; Somers, G. L.; Renfro, J. R. (2003) Ozone injury on cutleaf
18             coneflower (Rudbeckia laciniata) and crown-beard (Verbesina occidentalis) in Great Smoky Mountains
19             National Park. Environ. Pollut. 125: 53-60.
20       Chen, C. W.; Tsai, W. T.; Gomez, L. E. (1994) Modeling responses of Ponderosa pine to interacting stresses of ozone
21             and drought. For. Sci. 40: 267-288.
22       Chen, C. W.; Tsai, W. T.; Lucier,  A. A. (1998) A model of air-tree-soil system for ozone impact analysis. Ecol.
23             Modell. 111:207-222.
24       Cheong, Y. H.; Chang, H.-S.; Gupta, R.; Wang, X.; Zhu, T.; Luan, S.  (2002) Transcriptional profiling reveals novel
25             interactions between wounding, pathogen, abiotic stress, and hormonal responses in Arabidopsis. Plant
26             Physiol. 129: 661-677.
27       Chernikova, T.; Robinson, J. M.; Lee, E. H.; Mulchi, C. L. (2000) Ozone tolerance and antioxidant enzyme activity in
28             soybean cultivars. Photosynth. Res. 64: 15-26.
29       Chevrier, N.; Chung, Y. S.; Sarhan, F. (1990) Oxidative damages and repair inEuglena gracilis exposed to ozone: II.
3 0             membrane permeability and uptake of metabolites. Plant Cell Physiol. 31: 987-992.
31       Clark, C. S.; Weber, J. A.; Lee, E. H.; Hogsett, W. E. (1995)  Accentuation of gas exchange gradients in flushes of
32             ponderosa pine exposed to  ozone. Tree Physiol.  15:  181-189.
33       Clarke, B. B.; Henninger, M. R.; Brennan, E. (1983) An assessment of potato losses caused by oxidant air pollution in
34             New Jersey. Phytopathology 73:  104-108.
35       Cobb, F. W., Jr.; Wood, D. L.; Stark, R. W.; Parmeter, J. R., Jr. (1968) Photochemical oxidant injury and bark beetle
36             (Coleoptera: Scolytidae) infestation of ponderosa pine. IV. Theory on the relationships between oxidant injury
37             and bark beetle infestation. Hilgardia 39:  141-152.
38       Coe, H.; Gallagher, M. W.; Choularton, T. W.; Dore, C. (1995) Canopy scale measurements of stomatal and cuticular
39             O3 uptake by Sitka spruce. Atmos. Environ. 29: 1413-1423.
40       Coleman, J. S.; Jones, C. G.; Smith, W. H. (1987) The effect of ozone oncottonwood -leaf rust interactions:
41             independence of abiotic stress, genotype,  and leaf ontogeny. Can. J. Bot. 65: 949-953.
42       Coleman, M. D.; Dickson, R. E.; Isebrands, J. G.; Karnosky, D. F. (1995a) Carbon allocation and partitioning in
43             aspen clones varying in sensitivity to tropospheric ozone. Tree  Physiol. 15: 593-604.
44       Coleman, M. D.; Isebrands, J.  G.;  Dickson, R. E.; Karnosky, D. F. (1995b) Photosynthetic productivity of aspen
45             clones varying in sensitivity to tropospheric ozone. Tree Physiol. 15: 585-592.
46       Coleman, M. D.; Dickson, R. E.; Isebrands, J. G.; Karnosky, D. F. (1996) Root growth physiology of potted and
47             field-grown trembling aspen exposed to tropospheric ozone. Tree Physiol. 16: 145-152.
48       Colls, J. J.; Unsworth, M. H. (1992) Air pollution interactions with natural stressors. In: Barker, J. R.; Tingey, D. T.,
49             eds. Air pollution effects on biodiversity.  New York, NY: Van Nostrand Reinhold; pp. 93-108.
50       Common, M.; Perrings, C. (1992) Towards an ecological economics of sustainability. Ecol. Econ. 6: 7-34.
51       Constable, J. V. H.; Friend, A. L. (2000) Suitability of process-based tree growth models for addressing tree response
52             to climate change. Environ. Pollut. 110: 47-59.
53       Constable, J. V H.; Taylor, G. E., Jr. (1997) Modeling the  effects of elevated tropospheric O3 on two varieties of
54             Pinusponderosa. Can. J. For. Res. 27: 527-537.
55       Constable, J. V H.; Taylor, G. E., Jr.; Laurence, J. A.; Weber, J. A. (1996) Climatic change effects on the physiology
56             and growth of Pinus ponderosa: expectations from simulation modeling. Can. J. For. Res. 26: 1315-1325.
         January 2005                                    9-3 54        DRAFT-DO NOT QUOTE OR CITE

-------
  1       Cook, E. (1990) A conceptual linear aggregate model for tree rings. In: Cook, E. R.; Kairiukstis, L. A., eds. Methods
  2             of Dendrochronology: applications in the environmental sciences. Dordrecht, The Netherlands: Kluwer
  3             Academic Publishers, pp. 98-104.
  4       Cooley, D. R.; Manning, W. J. (1987) The impact of ozone on assimilate partitioning in plants: a review. Environ.
  5             PoUut. 47:95-113.
  6       Coruzzi, G. M.; Zhou, L. (2001) Carbon and nitrogen sensing and signaling in plants: emerging 'matrix effects'. Curr.
  7             Opin. Plant Biol. 4: 247-253.
  8       Costa, S. D.; Kennedy, G. G.; Heagle, A. S. (2001) Effect of host plant ozone stress on Colorado potato beetles.
  9             Environ. Entomol. 30: 824-831.
10       Costanza, R.; d'Arge, R.; De Groot, R.; Farber, S.; Grasso, M.; Harmon, B.; Limburg, K.; Naeem, S.; O'Neill, R. V.;
11             Paruelo, J.; Raskin, R. G.; Sutton, P.; Van Den Belt, M. (1997) The value of the world's ecosystem services
12             and natural capital.  Nature (London) 387: 253-259.
13
14       Coulston, J. W.; Smith, G. C.; Smith, W. D. (2003) Regional assessment of ozone sensitive tree species using
15             bioindicator plants. Environ. Monit. Assess. 83: 113-127.
16       Cox, R. M.; Malcolm, J. W. (1999) Passive ozone monitoring for forest health assessment. Water Air Soil Pollut. 116:
17             339-344.
18       Craigon, J.; Fangmeier, A.; Jones, M.; Donnelly, A.; Bindi, M.; De Temmerman, L.; Persson, K.; Ojanpera, K. (2002)
19             Growth and marketable-yield responses of potato to increased CO2 and ozone. Eur. J. Agron. 17: 273-289.
20       Crocker, T. D. (1985) On the value of the  condition of a forest stock. Land Econ. 61: 244-254.
21       Crocker, T. D.; Forster, B. A. (1986) Atmospheric deposition and forest decline. Water Air Soil Pollut. 31:
22             1007-1017.
23       Curtis, P. S. (1996) A meta-analysis of leaf gas exchange and nitrogen in trees grown under elevated carbon dioxide.
24             Plant Cell Environ. 19: 127-137.
25       Dahlsten, D. L.; Rowney, D. L.; Kickert, R. N. (1997) Effects of oxidant air pollutants on western pine beetle
26             (Coleoptera:  Scolytidae) populations in southern California. Environ. Pollut. 96: 415-423.
27       Daily, G. C. (1999) Developing a scientific basis for managing Earth's life support systems. Conserv. Ecol. 3: 14.
28             Available: http://www.ecologyandsociety.org/vol3/iss2/artl4/ [21 April, 2004].
29       Daily, G. C. (2000) Ecological assessment (untitled article in this section). Issues Sci. Technol. 16: 6.
30       Dammgen, U.; Griinhage,  L.; Haenel, H.-D.; Ja'ger, H.-J. (1993) Climate and stress in ecotoxicology. A coherent
31             system of definitions and terms. Angew. Bot. 67: 157-162.
32       Danielsson, H.; Gelang, J.; Pleijel, H. (1999) Ozone sensitivity, growth and flower development inPhleum genotypes
33             of different geographic origin in the Nordic countries. Environ. Exp. Bot. 42: 41-49.
34       Danielsson, H.; Karlsson,  G. P.; Karlsson, P. E.; Pleijel, J. (2003) Ozone uptake modelling and flux-response
3 5             relationships-an assessment of ozone-induced yield loss in spring wheat. Atmos. Environ. 37: 475-485.
3 6       Dass, H. C.; Weaver, G. M. (1968) Modification of ozone damage to phaseolus vulgaris by antioxidants, thiols and
37             sulfhydryl reagents. Can. J. Plant Sci. 48: 569-574.
38       Dass, H. C.; Weaver, G. M. (1972) Enzymatic changes in intact leaves of Phaseolus vulgaris following ozone
39             fumigation. Atmos. Environ. 6: 759-763.
40       Davidson, A. (1993) Update of ozone trends in California's South Coast Air Basin. Air Waste 43: 226-227.
41       Davis, D. D.; Skelly, J. M. (1992) Foliar sensitivity of eight eastern hardwood tree species to ozone. Water Air Soil
42             Pollut. 62: 269-277.
43       Davison, A. W.; Barnes, J. D. (1998) Effects of ozone on wild plants. New Phytol. 139: 135-151.
44       Davison, A. W.; Reiling, K. (1995) A rapid change in ozone resistance ofPlantago major after summers with high
45             ozone concentrations. New Phytol.  131: 337-344.
46       Davison, A. W.; Neufeld, H. S.; Chappelka, A. H.; Wolff, K. W.; Finkelstein, P. L. (2003) Interpreting spatial
47             variation in ozone symptoms shown by cutleaf cone flower, Rudbeckia laciniata L. Environ. Pollut. 125:
48             61-70.
49       De Kok, L. J.; Stulen, I. (1998) Responses of plant metabolism to air pollution and global change. Leiden, The
50             Netherlands: Backhuys Publishers.
51       De Santis, F. (1999) New directions: will a new European vegetation ozone standard be fair to all European
52             countries? Atmos. Environ. 33: 3873-3874.
53       De Santis, F. (2000) Reply to 'Use and abuse of the AOT40  concept' by M. Ashmore and J. Fuhrer [letter]. Atmos.
54             Environ. 34:  1158-1159.
55       De Temmerman, L.; Hacour, A.;  Guns, M. (2002a) Changing climate and potential impacts on potato yield and
56             quality 'CHIP': introduction, aims and methodology. Eur. J. Agron. 17: 233-242.
         January 2005                                   9-355        DRAFT-DO NOT QUOTE OR CITE

-------
  1       De Temmerman, L.; Wolf, I; Colls, I; Hindi, M; Fangmeier, A.; Finnan, J.; Ojanpera, K.; Pleijel, H. (2002b) Effect
  2             of climatic conditions on tuber yield (Solanum tuberosum L.) in the European 'CHIP' experiments. Eur. J.
  3             Agron. 17: 243-255.
  4       DeLucia, E. H.; Hamilton, J. G.; Naidu, S. L.; Thomas, R. B.; Andrews, J. A.; Finzi, A.; Lavine, M.; Matamala, R.;
  5             Mohan, J. E.; Hendrey, G. R.; Schlesinger, W. H. (1999) Net primary production of a forest ecosystem with
  6             experimental CO2 enrichment. Science (Washington, DC) 284: 1177-1179.
  7       Degousee, N.; Triantaphylides, C.; Montillet, J.-L. (1994) Involvement of oxidative processes in the signaling
  8             mechanisms leading to the activation of glyceollin synthesis in soybean (Glycine max). Plant Physiol. 104:
  9             945-952.
10       Devlin, W. S.; Gustine, D. L. (1992) Involvement of the oxidative burst in phytoalexin accumulation and the
11             hypersensitive reaction. Plant Physiol. 100: 1189-1195.
12       Diaz, G.; Barrantes, O.; Honrubia, M.; Gracia, C. (1996) Effect of ozone and sulphur dioxide on mycorrhizae of
13             Pinus halepensis Miller. Ann. Sci. For. 53: 849-856.
14       Dickson, R. E.; Coleman, M. D.; Reimenschneider, D. E.; Isebrands, J. G.; Hogan, G. D.; Karnosky, D. F. (1998)
15             Growth of five hybrid poplar genotypes exposed to interacting elevated CO2 and O3. Can. J. For. Res. 28:
16             1706-1716.
17       Dickson, R. E.; Coleman, M. D.; Pechter, P.; Karnosky, D. (2001) Growth and crown architecture of two aspen
18             genotypes exposed to interacting ozone and carbon dioxide. Environ. Pollut. 115: 319-334.
19       Dietz, K.-J. (1997) Functions and responses of the leaf apoplast under stress. Prog. Bot. 58: 221-254.
20       Dighton, J.; Jansen, A. E. (1991) Atmospheric pollutants and ectomycorrhizae: more questions than answers? Environ.
21             Pollut. 73: 179-204.
22       Dijak, M.; Ormrod, D. P. (1982) Some physiological and anatomical characteristics associated with differential ozone
23             sensitivity among pea cultivars. Environ. Exp. Bot. 22: 395-402.
24       Dixon, B. L.; Segerson, K. (1999) Impacts of increased climate variability on the profitability of Midwest agriculture.
25             J. Agric. Appl. Econ.  31: 537-548.
26       Dixon, M.; Le Thiec, D.; Garrec, J. P. (1998) Reactions of Norway spruce and beech trees to 2 years of ozone
27             exposure and episodic drought. Environ. Exp. Bot. 40: 77-91.
28       Docherty, M.; Salt, D. T.; Holopainen, J. K. (1997) The impacts of climate change and pollution on forest pests. In:
29             Watt, A. D.; Stork, N. E.; Hunter, M. D., eds. Forests and insects. New York, NY:  Chapman and Hall; pp.
30             229-247.
31       Dockery, D. W.; Schwartz, J.; Spengler, J. D. (1992) Air pollution and daily mortality: associations with particulates
32             and acid aerosols. Environ. Res. 59:  362-373.
33       Dominy, P. J.; Health, R. L. (1985) Inhibition of the K+-stimulated ATpase of the plasmalemma of pinto bean leaves
34             by ozone. Plant Physiol. 77: 43-45.
35       Donnelly, A.; Jones, M. B.; Schneiders, B.  J.; Burke, J.  I. (1998) The interactive effects of CO2, O3 and nitrogen on the
36             photosynthetic response of juvenile wheat plants. In: De Kok, L. J.; Stolen, L, eds.  Responses of plant
37             metabolism to air pollution and global change. Leiden, The Netherlands: Backhuys Publishers; pp. 291-293.
38       Donnelly, A.; Jones, M. B.; Burke, J. L; Schnieders, B.  (1999) Does elevated CO2 protect grain yield of wheat from the
39             effects of ozone stress? Z. Naturforsch. C: J. Biosci. 54: 802-811.
40       Donnelly, A.; Jones, M. B.; Burke, J. L; Schneiders, B.  (2000) Elevated CO2 provides protection from O3 induced
41             photosynthetic damage and chlorophyll loss in flag leaves of spring wheat (Triticum aestivum L., cv.'Minaret').
42             Agric. Ecosyst. Environ. 80: 159-168.
43       Donnelly, A.; Craigon, J.; Black, C. R.; Colls, J. J.; Landon, G. (2001a) Elevated CO2 increases biomass and tuber
44             yield in potato even at high ozone concentrations. New Phytol. 149: 265-274.
45       Donnelly, A.; Craigon, J.; Black, C. R.; Colls, J. J.; Landon, G. (2001) Does elevated CO2 ameliorate the impact of O3
46             on the chlorophyll content and photosynthesis in potato (Solanum tuberosum)! Physiol. Plant. 111:501-511.
47       Donnelly, A.; Lawson, T.; Craigon, J.; Black, C. R.; Coils, J. J.; Landon, G.  (200lb) Effects of elevated CO2 and O3 on
48             tuber quality in potato (Solanum tuberosum L.). Agric. Ecosyst. Environ. 87: 273-285.
49       Dowding, P. (1988) Air pollutant effects on plant pathogens. In: Schulte-Hostede, S.; Darrall, N. M.; Blank, L. W.;
50             Wellburn, A. R., eds.  Air pollution and plant metabolism. London, United Kingdom: Elsevier Applied
51             Scientific; pp. 329-355.
52       Drogoudi,  P. D.; Ashmore, M. R. (2000) Does elevated ozone have differing effects in flowering and deblossomed
53             strawberry? New Phytol. 147: 561-569.
54       Drogoudi,  P. D.; Ashmore, M. (2001) 14C-allocation of flowering and deblossomed strawberry in response to elevated
55             ozone. New Phytol. 152: 455-461.
         January 2005                                    9-356       DRAFT-DO NOT QUOTE OR CITE

-------
  1       Duchelle, S. F.; Skelly, J. M. (1981) Response of common milkweed to oxidant air pollution in the Shenandoah
  2             National Park in Virginia. Plant Dis. 65: 661-663.
  3       Duchelle, S. F.; Skelly, J. M.; Chevone, B. I. (1982) Oxidant effects on forest tree seedling growth in the Appalachian
  4             mountains. Water Air Soil Pollut. 18: 363-373.
  5       Duchelle, S. F.; Skelly, J. M.; Sharick, T. L.; Chevone, B. I.; Yang, Y.-S.; Nellessen, J. E. (1983) Effects of ozone on
  6             the productivity of natural vegetation in a high meadow of the Shenandoah National Park of Virginia. J.
  7             Environ. Manage. 17: 299-308.
  8       Dueck, Th. A.; Zuin, A.; Elderson, J. (1998) Influence of ammonia and ozone on growth and drought sensitivity of
  9             Pinus sylvestris. Atmos. Environ. 32: 545-550.
10       Dunn, D. B. (1959) Some effects of air pollution on Lupinus in the Los Angeles area. Ecology 40: 621-625.
11       Eamus, D. (1996) Responses of field grown trees to CO2 enrichment. Commonw. For. Rev. 75: 39-47.
12       Eamus, D.; Murray, M. (1991) Photo synthetic and stomatal conductance responses of Norway spruce and beech to
13             ozone, acid mist and frost - a conceptual model. Environ. Pollut. 72: 23-44.
14       Eckey-Kaltenbach, H.; Ernst, D.; Heller, W.; Sandermann, H., Jr. (1994) Biochemical responses to ozone. IV.
15             Cross-induction of defensive pathways in parsley (Petroselinum crispum L.) plants. Plant Physiol. 104: 67-74.
16       Eckey-Kaltenbach, H.; Grosskopf, E.; Sandermann, H., Jr.; Ernst, D. (1994) Induction of pathogen defence genes in
17             parsley (Petroselinum crispum  L.) plants by ozone. Proc. R. Soc. Edinburgh Sect. B: Biol. Sci. 102:  63-74.
18       Edwards, N. T. (1991) Root and soil respiration responses to ozone in Pinus taeda L. seedlings. New Phytol.  118:
19             315-321.
20       Edwards, G. S.; Kelly, J. M. (1992) Ectomycorrhizal colonization of loblolly pine seedlings during three growing
21             seasons  in response to ozone, acidic precipitation, and soil Mg status. Environ. Pollut. 76: 71-77.
22       Edwards, G. S.; Wullschleger, S. D.; Kelly, J. M. (1994) Growth and physiology of northern red oak: preliminary
23             comparisons of mature tree and seedling responses to ozone. Environ. Pollut.  83: 215-221.
24       Einig, W.; Lauxmann, U.; Hauch, B.;  Hampp, R.; Landolt, W.; Maurer, S.; Matyssek, R. (1997) Ozone-induced
25             accumulation of carbohydrates changes enzyme activities of carbohydrate metabolism in birch leaves. New
26             Phytol. 137: 673-680.
27       Elkiey, T.; Ormrod, D. P. (1979) Leaf diffusion resistance responses of three petunia cultivars to ozone and/or sulfur
28             dioxide. J. Air Pollut. Control Assoc. 29: 622-625.
29       Elkiey, T.; Ormrod, D. P. (1981) Sorption of O3, SO2, NO2 or their mixture by nine Poa pratensis cultivars  of differing
30             pollutant sensitivity. Atmos.  Environ. 15: 1739-1743.
31       Emberson, L. D.; Wieser, G.; Ashmore, M. R. (2000a) Modelling of stomatal conductance and ozone flux of Norway
32             spruce: comparison with field data. Environ. Pollut. 109: 393-402.
33       Emberson, L.; Ashmore, M. R.; Cambridge, H. M.; Simpson, D.; Tuovinen, J.-P. (2000b) Modelling stomatal ozone
34             flux across Europe.  Environ. Pollut. 109: 403-413.
3 5       Enders, G. (1992) Deposition of ozone to a mature spruce forest: measurements and comparison to models. Environ.
36             Pollut. 75: 61-67.
37       Ennis, C. A.; Lazrus, A. L.; Kok, G. L.; Zimmerman, P. R.; Monson, R. K. (1990) A branch chamber system and
3 8             techniques for simultaneous pollutant exposure experiments and gaseous flux determinations. Tellus Ser. B 42:
39             170-182.
40       Ernst, D.; Schraudner, M.; Langebartels, C.; Sandermann, H. (1992) Ozone-induced changes of messenger RNA levels
41             of beta-l,3-glucanase, chitinase and pathogenesis-related protein IB in tobacco plants. Plant. Mol. Biol. 20:
42             673-682.
43       Espelie, K. E.;  Franceschi, V. R.; Kolattukudy, P. E. (1986) Immunocytochemical localization and time course of
44             appearance of an anionic peroxidase associated with suberization in wound-healing potato tuber tissue. Plant
45             Physiol. 81:487-492.
46       Evans, P. A.; Ashmore, M. R. (1992) The effects of ambient air on a semi-natural grassland community. Agric.
47             Ecosyst. Environ. 38: 91-97.
48       Ewart, F.; Pleijel, H. (1999) Phenological development, leaf emergence, tillering and leaf area index, and duration of
49             spring wheat across Europe in response to CO2 and ozone. Eur.  J. Agron. 10:  171-184.
50       Ewart, F.; VanOijen, M.; Porter, J. R. (1999) Simulation of growth and development processes of spring wheat in
51             response to CO2 and ozone for  different sites and years in Europe using mechanistic crop simulation models.
52             Eur. J. Agron. 10: 231-247.
53       Fangmeier, A.; Stein, W.; Ja'ger, H.-J. (1992) Advantages of an open-top chamber plant exposure system to assess the
54             impact of atmospheric trace gases on vegetation. Angew. Bot. 66: 97-105.
         January 2005                                    9-357        DRAFT-DO NOT QUOTE OR CITE

-------
  1       Fangmeier, A.; Brockerhoff, U.; Graters, U.; Jager, H.-J. (1994a) Growth and yield responses of spring wheat
  2             (Triticum aestivum L. cv. Turbo) grown in open-top chambers to ozone and water stress. Environ. Pollut. 83:
  3             317-325.
  4       Fangmeier, A.; Branschon, S.; Jager, H.-J. (1994b) Time course of oxidant stress biomarkers in flag leaves of wheat
  5             exposed to ozone and drought stress. New Phytol. 126: 63-69.
  6       Fangmeier, A.; Graters, U.; Vermehren, B.; Jager, H.-J. (1996) Responses of some cereal cultivars to CO2 enrichment
  7             and tropospheric ozone at different levels of nitrogen supply. J. Appl. Bot. 70: 12-18.
  8       Fangmeier, A.; De Temmerman, L.; Mortensen, L.; Kemp, K.; Burke, J.; Mitchell, R.; van Oijen, M; Weigel, H. J.
  9             (1999) Effects on nutrients and on grain quality in spring wheat crops grown under elevated CO2 concentrations
10             and stress conditions in the European, multiple-site experiment 'ESPACE-wheat'. Eur. J. Agron. 10: 215-229.
11       Fangmeier, A.; Bender, J.; Weigel,  H.-J.; Jager, H.-J. (2002) Effects of pollutant mixtures. In: Bell, J. N. B.; Treshow,
12             M., eds. Air Pollution and Plant Life. 2nd ed. Chichester, United Kingdom:  John Wiley & Sons Ltd.; pp.
13             251-272.
14       Fangmeier, A.; De Temmerman, L.; Black, C.; Persson, K.; Vorne, V. (2002) Effects of elevated CO2 and/or ozone  on
15             nutrient concentrations and nutrient uptake of potatoes. Eur. J. Agron.  17: 353-368.
16       Farage, P. K.; Long, S. P. (1999) The effects of O3 fumigation during leaf development on photosynthesis of wheat  and
17             pea:  an in vivo analysis. Photosynth. Res. 59: 1-7.
18       Farage, P. K.; Long, S. P.; Lechner, E. G.; Baker, N. R. (1991) The sequence of changes within the photosynthetic
19             apparatus of wheat following short-term exposure to ozone.  Plant Physiol. 95: 529-535.
20       Farquhar, G. D.; Caemmerer, S. V.; Berry, J. A. (1980) A biochemical model of photosynthetic CO2 assimilation in
21             leaves of C3 species. Planta 149: 78-90.
22       Farquhar, G. D.; Sharkey, T. D. (1982) Stomatal conductance and photosynthesis. Annu. Rev. Plant Physiol. 33:
23             317-345.
24       Federal Land Manager's Air Quality Related Values Workgroup (FLAG). (2000) Phase I report. Denver, CO: National
25             Park Service, Air Resources Division. Available: http://www2.nature.nps.gov/ard/flagfree/ [11 March, 2003].
26       Federal Register. (1997) National ambient air quality standards for ozone; final rale. F. R. (July 18) 62: 38856-38896.
27       Fenn, M. E.; Bytnerowicz, A. (1993) Dry deposition of nitrogen and sulfur to ponderosa and Jeffrey pine in the San
28             Bernardino National Forest in southern California. Environ.  Pollut. 81: 277-285.
29       Fenn, M. E.; Dunn, P. H. (1989) Litter decomposition across an air-pollution gradient in the San Bernardino
30             Mountains. Soil Sci. Soc. Am. J. 53: 1560-1567.
31       Fenn, M. E.; Poth, M. A. (1998) Indicators of nitrogen status in California forests. U.S.D.A. Forest Service Gen. Tech.
32             Rep. PSW-GTR-166.
33       Fenn, M. E.; Dunn, R. H.; Wilborn, R. (1990) Black stain root disease in ozone-stressed ponderosa pine. Plant Dis.  74:
34             426-430.
35       Fenn, M. E.; Poth, M. A.; Johnson, D. W. (1996) Evidence for nitrogen saturation in the San Bernardino Mountains in
36             southern California. For. Ecol. Manage. 82: 211-230.
37       Fenn, M. E.; Poth, M. A.; Aber, J. D.; Baron, J. S.; Bormann, B. T.; Johnson, D. W.; Lemly,  A. D.; McNulty, S. G.;
3 8             Ryan, D. F.; Stottlemyer, R.  (1998) Nitrogen excess in North American ecosystems: predisposing factors,
39             ecosystem responses, and management strategies. Ecol. Appl. 8: 706-733.
40       Fenn, M. E.; Poth, M. A.; Schilling, S. L.; Grainger, D. B. (2000) Throughfall and fog deposition of nitrogen and
41             sulfur at an N-limited and N-saturated site in the San Bernardino Mountains, southern California. Can. J. For.
42             Res. 30: 1476-1488.
43       Fenn, M. E.; Poth, M. A.; Bytnerowicz, A.; Sickman, J. O.; Takemoto, B. K.  (2003) Effects of ozone, nitrogen
44             deposition, and other stressors on montane ecosystems in the Sierra Nevada. In: Bytnerowicz, A.; Arbaugh, M.
45             J.; Alonso, R., eds. Ozone air pollution in the Sierra Nevada: distribution and effects on forests; pp. 111-155.
46             New York, NY: Elsevier. (Developments in Environmental Science: v. 2).
47       Fenn, M. E.; Haeuber, R.; Tonnesen, G. S.; Baron, J. S.; Grossman-Clarke, S.; Hope, D.; Jaffe, D. A.;  Copeland, S.;
48             Geiser, L.; Rueth,  H. M.; Sickman, J. O. (2003a) Nitrogen emissions, deposition, and monitoring in the western
49             United States. BioScience 53: 391-403.
50       Fenn, M. E.; Baron,  J. S.; Allen, E.  B.; Rueth, H. M.; Nydick, K. R.; Geiser, L.; Bowman, W. D.; Sickman, J. O.;
51             Meixner, T.; Johnson, D. W.; Neitlich, P. (2003b) Ecological effects of nitrogen deposition in the western
52             United States. BioScience 53: 404-420.
53       Ferdinand, J. A.; Fredericksen, T. S.; Kouteric, K. B.; Skelly, J. M. (2000) Leaf morphology  and ozone sensitivity of
54             two open pollinated genotypes of black cherry (Primus serotina) seedlings. Environ. Pollut. 108: 297-302.
55       Findlay, S.; Jones, C. G. (1990) Exposure of cottonwood plants to ozone alters subsequent decomposition. Oecologia
56             82:248-250.
         January 2005                                    9-358       DRAFT-DO NOT QUOTE OR CITE

-------
  1       Findlay, S.; Carreiro, M.; Krischik, V.; Jones, C. G. (1996) Effects of damage to living plants on leaf litter quality.
  2             Ecol. Appl. 6: 269-275.
  3       Findley, D. A.; Keever, G. I; Chappelka, A. H.; Bakes, D. I; Gillian, D. J. (1997) Differential responses of buddleia
  4             (Buddleia davidii Franch.) to ozone. Environ. Pollut. 98: 105-111.
  5       Finkelstein, R. R.; Gibson, S. I. (2002) ABA and sugar interactions regulating development: cross-talk or voices in a
  6             crowd? Curr. Opin. Plant Biol. 5: 26-32.
  7       Finnan, J. M.; Jones, M. B.; Burke, J. I. (1996) A time-concentration study on the effects of ozone on spring wheat
  8             (Triticum aestivum L.). 2. A comparison of indices. Agric. Ecosyst. Environ. 57: 169-177.
  9       Finnan, J. M.; Burke, J. L.; Jones, M. B. (1997) An evaluation of indices that describe the impact of ozone on the yield
10             of spring wheat (Triticum aestivum L.) Atmos. Environ. 31:  2685-2693.
11       Finnan, J. M.; Donnelly, A.; Burke, J. L; Jones,  M. B. (2002) The effects of elevated concentrations of carbon dioxide
12             and ozone on potato (Solarium tuberosum L.) yield. Agric. Ecosyst. Environ. 88: 11-22.
13       Fiscus, E. L.; Reid, C. D.; Miller, J. E.; Heagle,  A. S. (1997) Elevated CO2 reduces O3 flux and O3-induced yield losses
14             in soybeans: possible implications for elevated CO2 studies.  J. Exp. Bot. 48: 307-313.
15       Fishman, J.; Watson, C. E.; Larsen, J. C.; Logan, J. A. (1990) Distribution of tropospheric ozone determined from
16             satellite data. J. Geophys. Res. [Atmos.] 95: 3599-3617.
17       Flagler, R. B., ed. (1998) Recognition of air pollution injury to vegetation: a pictorial atlas. 2nd ed. Pittsburgh, PA: Air
18             & Waste Management Association.
19       Fluckiger, W.; Braun, S.; Hiltbrunner, E. (2002) Effects of air pollutants onbiotic stress. In: Bell, J. N. B.; Treshow,
20             M., eds. Air Pollution and Plant Life. 2nd ed. Chichester, United Kingdom:  John Wiley & Sons Ltd.; pp.
21             379-406.
22       Fontan, J. A.; Minga, A.; Lopez, A.; Druilhet, A. (1992) Vertical ozone profiles in a pine forest. Atmos. Environ. Part
23             A 26A: 863-869.
24       Foot, J. P.; Caporn, S. J. M.;  Lee, J. A.; Ashenden, T. W. (1996) The effect of long-term ozone fumigation on the
25             growth, physiology and frost sensitivity of Calluna vulgaris. New Phytol. 133: 505-511.
26       Foot, J. P.; Caporn, S. J. M.;  Lee, J. A.; Ashenden, T. W. (1997) Evidence that ozone exposure increases the
27             susceptibility of plants to natural frosting episodes. New Phytol.  135: 369-374.
28       Fortin, M.; Maufette, Y.; Albert, P. J. (1997) The effects of ozone-exposed sugar maple seedlings on the biological
29             performance and the feeding preference of the forest tent catepillar (Malacosoma disstria Hbn.). Environ.
30             Pollut. 97: 303-309.
31       Fowler, D.; Cape, J. N. (1982) Air pollutants in agriculture and horticulture. In: Unsworth, M. H.; Ormrod, D. P.
32             Effects of gaseous air pollution in agriculture and horticulture. London, United Kingdom: Butterworth
33             Scientific; pp. 3-26.
34       Fox, S.; Mickler, R. A., eds. (1996) Impact of air pollutants on southern pine forests. New York, NY: Springer-Verlag.
35             (Lange, O. L.; Mooney, H. A., eds. Ecological studies: v. 118).
36       Foyer, C. H.; Souriau, N.; Ferret, S.; Lelandais,  M.; Kunert, K.-J.; Pruvost, C.;  Jouanin, L. (1995) Overexpression of
37             glutathione reductase but not glutathione synthetase leads to increases in antioxidant capacity and resistance to
3 8             photoinhibition in poplar trees. Plant Physiol. 109:  1047-1057.
39       Frederick, P. E.; Heath, R. L. (1970) Ozone-induced fatty acid and viability changes in Chlorella. Plant Physiol. 55:
40             15-19.
41       Fredericksen, T. S.; Joyce, B. J.; Skelly, J. M.; Steiner, K. C.; Kolb, T. E.; Kouterick, K. B.; Savage, J. E.; Snyder, K.
42             R. (1995) Physiology, morphology, and ozone uptake of leaves of black cherry seedlings, saplings, and canopy
43             trees. Environ. Pollut. 89: 273-283.
44       Fredericksen, T. S.; Kolb, T. E.; Skelly, J. M.; Steiner, K. C.; Joyce, B.  J.; Savage, J. E. (1996a) Light environment
45             alters ozone uptake per net photosynthetic rate in black cherry trees. Tree Physiol. 16: 485-490.
46       Fredericksen, T. S.; Skelly, J. M.; Snyder, K. R.; Steiner, K. C.; Kolb, T. E. (1996b) Predicting ozone uptake from
47             meteorological and environmental variables. J. Air Waste Manage. Assoc. 46: 464-469.
48       Freer-Smith, P. H.; Taylor, G. (1992) Comparative evaluation of the effects of gaseous pollutants, acidic deposition
49             and mineral deficiencies on gas exchange of trees. Agric. Ecosyst. Environ. 42: 321-332.
50       Froehlich, J. E.; Itoh, A.; Howe, G. A. (2001) Tomato allene oxide synthase and fatty acid hydroperoxide lyase, two
51             cytochrome P450s involved in oxylipin metabolism, are targeted to different membranes of chloroplast
52             envelope. Plant Physiol. 125:  306-317.
53       Fuentes, J. D.; Gillespie, T. J. (1992) A gas exchange system to study the effects of leaf surface wetness on the
54             deposition of ozone. Atmos. Environ. Part A 26: 1165-1173.
55       Fuentes, J. D.; Gillespie, T. J.; den Hartog, G.; Neumann, H. H. (1992) Ozone deposition onto a deciduous forest
56             during dry and wet conditions. Agric. For. Meteorol. 62: 1-18.
         January 2005                                     9-359        DRAFT-DO NOT QUOTE OR CITE

-------
  1       Fuentes, J. D.; Hayden, B. P.; Garstang, M.; Lerdau, M; Fitzjarrald, D.; Baldocchi, D. D.; Monson, R.; Lamb, B.;
  2             Geron, C. (2001) New directions:  VOCs and biosphere-atmosphere feedbacks. Atmos. Environ. 35: 189-191.
  3       Fuhrer, J. (1994) Effects of ozone on managed pasture. 1. Effects of open-top chambers on microclimate, ozone flux,
  4             and plant growth. Environ. Pollut. 86: 297-305.
  5       Fuhrer, J. (1994) The critical level for ozone to protect agricutural crops - an assessment of data from European
  6             open-top chamber experiments. In: Fuhrer, J.; Achermann, eds. Critical levels for ozone: a UN-ECE workshop
  7             report; FAC report no. 16. Liebefeld-Bern, Switzerland: Swiss Federal Research Station for Agricultural
  8             Chemistry and Environmental Hygiene; pp. 42-57.
  9       Fuhrer, J. (1996) The critical level for effects of ozone on crops and the transfer to mapping. In: Karenlampi, L.;
10             Skarby, L., eds. Critical levels for ozone in Europe: testing and finalizing the concepts: UN-ECE workshop
11             report. UN-ECE Convention on Long-Range Transboundary Air Pollution; April; Kuopio, Finland. Kuopio,
12             Finland: University of Kuopio, Department of Ecology and Environmental Science; pp. 27-43.
13       Fuhrer, J.; Grimm, A. G.; Tschannen, W.; Shariat-Madari, H. (1992) The response of spring wheat (Triticum aestivum
14             L.) to ozone at higher elevations: II. changes in yield, yield components and grain quality in response to ozone
15             flux. New Phytol. 121: 211-219.
16       Fuhrer, J.; Shariat-Madari, H.; Perler, R.; Tschannen, W.; Grub, A. (1994) Effects of ozone on managed pasture: II.
17             Yield, species composition, canopy structure, and forage quality. Environ. Pollut. 86: 307-314.
18       Fuhrer,!.; Skarby, L.; Ashmore, M. R. (1997) Critical levels for ozone effects on vegetation in Europe. Environ.
19             Pollution 97: 91-106.
20       Fumagalli, L; Mignanego, L.; Violini, G. (1997) Effects of tropospheric ozone on white clover plants exposed in
21             open-top chambers or protected by the antioxidant ethylene-diurea (EDU). Agronomic 17: 271-281.
22       Fumagalli, L; Ambrogi, R.; Mignanego, L. (2001) Yield responses of plants exposed to ambient ozone in the river Po
23             valley (Italy). Agronomic 21: 227-233.
24       Garcia, P.; Dixon, B.; Mjelde, J.; Adams, R. (1986) Measuring the benefits of environmental change using a duality
25             approach: the case of ozone and Illinois cash grain farms. J. Environ. Econ. Manage. 13: 69-80.
26       Gasper, T.; Penel, C.; Castillo, F. J.; Greppin, H. (1985) A two step control of basic and acidic peroxidases and its
27             significance for growth and development. Physiol. Plant. 64: 418-423.
28       Gate, I. M.; McNeill, S.; Ashmore, M. R. (1995) Effects of air pollution on the searching behaviour of an insect
29             parasitoid. Water Air Soil Pollut. 85: 1425-1430.
30       Gates, D. M. (1968) Transpiration and leaf temperature. Annu. Rev. Plant Physiol. 19: 211-238.
31       Gaucher, C.; Costanzo, N.; Afif, D.; Mauffette, Y.; Chevrier, N.; Dizengremel, P. (2003) The impact of elevated ozone
32             and carbon dioxide on young^cer saccharum seedlings. Physiol. Plant. 117: 392-402.
33       Gelang, J.;Pleijel, H.; Sild, E.; Danielsson, H.; Younis, S.; Sellden, G. (2000) Rate and duration of grain filling in
34             relation to flag leaf senescence and grain yield in spring wheat (Triticum aestivum) exposed to different
35             concentrations of ozone.  Physiol. Plant. 110: 366-375.
36       Gerant, D.; Podor, M.; Grieu, P.; Afif, D.; Cornu,  S.; Morabito, D.; Banvoy, J.; Robin, C.; Dizengremel, P. (1996)
3 7             Carbon metabolism enzyme activities and carbon partitioning in Pinus halepensis mill. Exposed to mild drought
38             and ozone.  J. Plant Physiol. 148: 142-147.
39       Gerosa, G.; Marzuoli, R.; Bussotti, F.; Pancrazi, M.; Ballarin-Denti, A. (2003) Ozone sensitivity ofFagus sylvatica
40             and Fraxinus excelsior young trees in relation to leaf structure and foliar ozone uptake. Environ. Pollut. 125:
41             91-98.
42       Ghosh, S.; Skelly, J. M.; Innex, J. L.; Skelly, L. (1998) Temporal development of visual ozone injury on the foliage of
43             Primus serotina - a statistical evaluation. Environ. Pollut. 192: 287-300.
44       Gifford, R. M.; Morison, J. I. L. (1993) Crop responses to the  global increase in atmospheric carbon dioxide
45             concentration. In: Buxton, D. R., et al., eds. International crop science I. Madison, WI: Crop Science Society of
46             America; pp. 325-331.
47       Gimeno, B. S.; Bermejo, V.; Reinert, R. A.; Zheng, Y.; Barnes, J. D. (1999) Adverse effects of ambient ozone on
48             watermelon yield and physiology at a rural site  in eastern Spain. New Phytol.  144: 245-260.
49       Goldstein, A. H.; Kurpius, M. R.; Panek, J. A. (2003) Ozone uptake by ponderosa pine in the Sierra Nevada: a
50             measurement perspective. In: Bytnerowicz, A.;  Arbaugh, M. J.; Alonso, R., eds. Ozone air pollution in the
51             Sierra Nevada: distribution and effects on forests; pp. 83-109. New York, NY: Elsevier. (Developments in
52             Environmental Science: v. 2).
53       Goodyear, S. N.; Ormrod, D. P. (1988) Tomato response to concurrent and sequential NO2 and O3 exposures.  Environ.
54             Pollut. 51: 315-326.
         January 2005                                    9-360        DRAFT-DO NOT QUOTE OR CITE

-------
  1       Grains, T. E. E.; Anegg, S.; Haberle, K.-H.; Langebartels, C.; Matyssek, R. (1999) Interactions of chronic exposure to
  2             elevated CO2 and O3 levels in the phytosynthetic light and dark reactions of European beech (Fagus sylvatica).
  3             New Phytol. 144: 95-107.
  4       Grantz, D. A.; Farrar, J. F. (1999) Acute exposure to ozone inhibits rapid carbon translocation from source leaves of
  5             Pima cotton. J. Exp. Bot. 50: 1253-1262.
  6       Grantz, D. A.; Yang, S. D. (1996) Effect of O3 on hydraulic architecture in pima cotton. Plant Physiol. 112: 1649-1657.
  7       Grantz, D. A.; Yang, S. (2000) Ozone impacts on allometry and root hydraulic conductance are not mediated by source
  8             limitation nor developmental age. J. Exp. Bot. 51: 919-927.
  9       Grantz, D. A.; Zhang, X. J.; Massman, W. J.; DenHartog, G.; Neumann, H. H.; Pederson, J. R. (1995) Effects of
10             stomatal conductance and surface wetness on ozone deposition in field-grown grape. Atmos. Environ. 29:
11             3189-3198.
12       Grantz, D. A.; Zhang, X. J.; Massman, W.; Delany, A.; Pederson, R. (1997) Ozone deposition to a cotton (Gossypium
13             hirsutum L.) field: stomatal and surface wetness effects during the California Ozone Deposition experiment.
14       Agric. For. Meteorol. 85: 19-31.
15       Graumlich, L. J. (1993) A 1000-year record of temperature and precipitation in the Sierra Nevada. Quaternary Res. 39:
16             249-255.
17       Greenwood, P.; Greenhalgh, A.; Baker, C.; Unsworth, M. (1982) A computer-controlled system for exposing field
18             crops to gaseous air pollutants. Atmos. Environ. 16: 2261-2266.
19       Gregg, J. W.; Jones, C. G.; Dawson, T.  E. (2003) Urbanization  effects on tree growth in the vicinity of New York City
20             [letter]. Nature (London, U.K.) 424: 183-187.
21       Greitner, C. S.; Pell, E. J.; Winner, W. E. (1994) Analysis of aspen foliage exposed to multiple stresses: ozone,
22             nitrogen deficiency and drought. New Phytol. 127: 579-589.
23       Grimes, H. D.; Perkins, K. K.; Boss, W. F. (1983) Ozone degrades into hydroxyl radical under physiological
24             conditions. Plant Physiol. 72: 1016-1020.
25       Grosjean, D.; Williams, E. L., II; Grosjean, E. (1995) Monitoring ambient  ozone with a network of passive samplers: a
26             feasibility study. Environ. Pollut. 88: 267-273.
27       Gross, G. G.; Janse, C. (1977) Formation of NADH and hydrogen peroxide by cell wall-associated enzymes from
28             forsythia xylem. Z. Pflanzenphysiol. 84: 447-452.
29       Groth, J. V.; Krupa, S. V. (2000) Crop ecosystem responses to climatic change: interactive effects of ozone,
30             ultraviolet-B radiation, sulphur dioxide and carbon dioxide on crops. In: Reddy, K.  R.; Hodges, H. F., eds.
31             Climate change and global crop productivity. New York, NY: CABI Publishing; pp. 387-405.
32       Grulke, N. E. (1999) Physiological responses of ponderosa pine to gradients of environmental stressors. In: Miller, P.
33             R.; McBride, J. R., eds. Oxidant  Air Pollution Impacts in the Montane Forests of Southern California. New
34             York, NY: Springer-Verlag; pp.  126-163. (Ecological Studies, v. 134).
3 5       Grulke, N. E.; Balduman, L. (1999) Deciduous conifers: high N deposition and O3 exposure effects on growth and
36             bio mass allocation in ponderosa pine. Water Air Soil Pollut. 116: 235-248.
37       Grulke, N. E.; Lee, E. H. (1997) Assessing visible ozone-induced injury in ponderosa pine. Can. J. For. Res. 27:
38             1658-1668.
39       Grulke, N. E.; Miller, P. R. (1994) Changes in gas-exchange characteristics during the life-span of giant sequoia:
40             implications for response to current and future concentrations of atmospheric ozone. Tree Physiol. 14: 659-668.
41       Grulke, N. E.; Retzlaff, W. A. (2001) Changes in physiological attributes of ponderosa pine from seedling to mature
42             tree. Tree Physiol. 21: 275-286.
43       Grulke, N. E.; Miller, P. R.; Scioli, D. (1996) Response of giant sequoia canopy foliage to elevated concentrations of
44             atmospheric ozone. Tree Physiol. 16: 575-581.
45       Grulke, N. E.; Andersen, C. P.; Fenn, M. E.; Miller, P. R. (1998) Ozone exposure and nitrogen deposition lowers root
46             biomass of ponderosa pine in the San Bernardino Mountains, California. Environ. Pollut. 103: 63-73.
47       Grulke, N. E.; Miller, P. R.; Leininger, T. D. (1998) Effect of ozone exposure on seasonal gas exchange of five
48             western conifers. In: Bytnerowicz, A.; Arbaugh, M. J.; Schilling, S. L., eds. Proceedings of the International
49             Symposium on Air Pollution and Climate Change Effects on Forest Ecosystems; February, 1996; Riverside,
50             CA. Albany, CA: USDA Forest Service, Pacific Southwest General Technical Report no. PSW-GTR 166; pp.
51             229-238.
52       Grulke, N. E.; Andersen, C. P.; Hogsett, W. E. (2001) Seasonal changes in above- and belowground carbohydrate
53             concentrations of ponderosa pine along a pollution gradient. Tree Physiol. 21: 173-181.
         January 2005                                    9-361        DRAFT-DO NOT QUOTE OR CITE

-------
  1       Grulke, N. E.; Preisler, H. K.; Fan, C. C.; Retzlaff, W. A. (2002) A statistical approach to estimate O3 uptake of
  2             ponderosa pine in a Mediterranean climate. Environ. Pollut. 119: 163-175.
  3       Grulke, N. E.; Preisler, H. K.; Rose, C.; Kirsch, J.; Balduman, L. (2002) O3 uptake and drought stress effects on carbon
  4             acquisition of ponderosa pine in natural stands. New Phytol. 154: 621-631.
  5       Grulke, N. E.; Johnson, R.; Esperanza, A.; Jones, D.; Nguyen, T.; Posch, S.; Tausz, M. (2003a) Canopy transpiration
  6             of Jeffrey pine in mesic and xeric microsites:  O3 uptake and injury response. Trees (Berlin, Ger.) 17: 292-298.
  7       Grulke, N. E.; Johnson, R.; Monschein, S.; Nikolova, P.; Tausz, M. (2003b) Variation in morphological and
  8             biochemical O3 injury attributes of mature Jeffrey pine within canopies and between microsites. Tree Physiol.
  9             23: 923-929.
10       Grulke, N. E.; Alonso, R.; Nguyen, T.; Cascio, C.; Dobrowolski, W. (2004) Stomata open at night in pole-sized and
11             mature ponderosa pine: implications for O3 exposure metrics. Tree Physiol. 24: 1001-1010.
12       Griinhage, L.; Haenel, H.-D. (1997) PLATIN (PLant-ATmosphere-INteraction) I: a model of plant-atmosphere
13             interaction for estimating absorbed doses of gaseous air pollutants. Environ. Pollut. 98: 37-50.
14       Griinhage, L.; Jager, H.-J. (1994) Influence of the atmospheric conductivity on the ozone exposure of plants under
15             ambient conditions: considerations for establishing ozone standards to protect vegetation. Environ. Pollut. 85:
16             125-129.
17       Griinhage, L.; Jager, H.-J. (1996) Critical levels for ozone, ozone exposure potentials of the atmosphere or critical
18             absorbed doses for ozone: a general discussion. In: Karenlampi, L.;  Skarby, L., eds. Critical levels for ozone in
19             Europe: testing and finalizing the concepts: UN-ECE workshop report. UN-ECE Convention on Long-Range
20             Transboundary Air Pollution; April; Kuopio,  Finland. Kuopio, Finland: University of Kuopio, Department of
21             Ecology and Environmental Science; pp. 151-168.
22       Griinhage, L.; Jager, H.-J. (2003) From critical levels to critical loads for ozone: a discussion of a new experimental
23             and modelling approach for establishing flux-response relationships for agricultural crops and native plant
24             species. Environ. Pollut. 125: 99-110.
25       Griinhage, L.; Dammgen, U.; Haenel, H.-D.; Jager, H.-J.; Holl,  A.; Schmitt, J.; Hanewald, K. (1993) A new potential
26             air-quality criterion derived from vertical flux densities of ozone and from plant response.  Angew. Bot. 67:
27             9-13.
28       Griinhage, L.; Dammgen, U.; Haenel, H.-D.; Jager, H.-J. (1994) Response  of a grassland ecosystem to air pollutants:
29             III~the chemical climate: vertical flux densities of gaseous species in the atmosphere near the ground. Environ.
30             Pollut. 85: 43-49.
31       Griinhage, L.; Jager, H.-J.; Haenel, H.-D.; Hanewald, K.; Krupa, S. (1997) PLATIN (plant-atmosphere interaction) II:
32             co-occurrence of high ambient ozone concentrations and factors limiting plant absorbed dose. Environ. Pollut.
33             98:51-60.
34       Griinhage, L.; Jager, H.-J.; Haenel, H.-D.; Lopmeier, F.-J.; Hanewald, K. (1999) The European critical levels for
35             ozone: improving their usage. Environ. Pollut. 105: 163-173.
36       Griinhage, L.; Krause, G. H. M.; Kollner, B.; Bender, J.; Weigel, H.-J; Jager, H.-J.; Guderian, R.  (2001)  A new
37             flux-orientated concept to derive critical levels for ozone to protect vegetation. Environ. Pollut.  Ill: 355-362.
38       Griinhage, L.; Krupa, S. V.; Legge, A. H.; Jager, H.-J. (2004) Ambient flux-based critical values  of ozone for
39             protecting vegetation: differing spatial scales  and uncertainties in risk assessment. Atmos.  Environ. 38:
40             2433-2437.
41       Grunwald, C.; Endress, A. G. (1985) Foliar sterols in soybeans exposed to  chronic levels of ozone. Plant Physiol. 77:
42             245-247.
43       Graters, U.; Fangmeier, A.;  Jager, H.-J. (1995) Modelling stomatal responses of spring wheat (Triticum aestivum L.
44             cv. Turbo) to ozone and different levels of water supply.  Environ. Pollut. 87: 141-149.
45       Guderian, R. (1977) Discussion  of the suitability of plant responses as a basis for air pollution control measures. In:
46             Air pollution: phytotoxicity of acidic gases and its significance  in air pollution control. Berlin, Federal Republic
47             of Germany: Springer Verlag; pp. 75-97. (Billings, W. D.; Golley, F.; Lange, O. L.; Olson, J. S., eds. Ecological
48             studies: analysis and  synthesis, v. 22).
49       Guidi, L.; Tonini, M.; Soldatini, G. F. (2000) Effects of high light and ozone fumigation on photosynthesis in
50             Phaseolus vulgaris. Plant Physiol. Biochem. 38: 717-725.
51       Guzy, M. R.; Heath R. L. (1993) Responses to ozone of varieties of common bean (Phaseolus vulgaris L.) New
52             Phytol. 124: 617-625.
         January 2005                                     9-362        DRAFT-DO NOT QUOTE OR CITE

-------
  1       Haber, F. (1924) Zur Geschichte des Gaskrieges [On the history of the gas war]. In: Fuenf Vortraege ausden Jahren
  2             1920-1923 [Five lectures from the years 1920-1923]. Berlin, Germany: Verlag von Julius Springer; pp. 76-92.
  3       Hacour, A.; Craigon, I; Vandermeiren, K.; Ojanpera, K.; Pleijel, H.; Danielsson, H.; Hogy, P.; Finnan, I; Bindi, M.
  4             (2002) CO2 and ozone effects on canopy development of potato crops across Europe. Eur. J. Agron. 17:
  5             257-272.
  6       Haberle, K. H.; Werner, H.; Fabian, P.; Pretzch, H.; Reiter, I.; Matyssek, R.  (1999) 'Free-air' ozone fumigation of
  7             mature forest trees: a concept for validitating AOT40 under stand conditions. In: Fuhrer, J.; Achermann, B., eds.
  8             Critical levels for ozone: level II: workshop under the Convention on Long-Range Transboundary Air Pollution
  9             of the United Nations Economic Commission for Europe (UN/ECE);  April; Gerzensee, Switzerland. Berne,
10             Switzerland: Bundesamt fur Umwelt, Wald und Landschaft; pp. 133-137.
11       Hakala,  K.; Mela, T. (1996) The effects of prolonged exposure to elevated temperatures and elevated CO2 levels on the
12             growth, yield and dry matter partitioning of field-sown meadow fescue. Agric. Food Sci. Finl. 5: 285-298.
13       Hankamer, B.; Barber, J.; Boekema, E. J. (1997) Structure and membrane organization of photo system II  in green
14             plants. Ann. Rev. Plant Physiol. Mol. Biol.  48: 641-671.
15       Hanson, P. J.; Samuelson, L. J.; Wullschleger, S. D.; Tabberer,  T. A.; Edwards, G.  S. (1994) Seasonal patterns of
16             light-saturated photosynthesis and leaf conductance for mature and seedling Quercus rubra L. foliage:
17             differential  sensitivity to ozone exposure. Tree Physiol. 14: 1351-1366.
18       Hao, X.; Hale, B. A.; Ormrod, D. P.; Papadopoulos, A. P. (2000) Effects of pre-exposure to ultraviolet-B  radiation on
19             responses of tomato (Lycopersicon esculentum cv. New Yorker) to ozone in ambient and elevated carbon
20             dioxide. Environ. Pollut. 110: 217-224.
21       Harris, M. J.; Bailey-Serres, J. N. (1994) Ozone effects on gene expression and molecular approaches to breeding for
22             air pollution resistance. In: Basra,  A.  S., ed. Stress-induced gene expression in plants. Langhorne, PA: Harwood
23             Academic Publishers; pp. 185-207.
24       Harward, M.; Treshow, M. (1975) Impact of ozone on the growth and reproduction of understorey plants  in the Aspen
25             zone of western U.S.A. Environ. Conserv. 2: 17-23.
26       Hausladen, A.; Madamanchi, N. R.; Fellows, S.; Alscher, R. G.; Amundson, R. G. (1990) Seasonal changes in
27             antioxidants in red spruce as affected by ozone. New Phytol.  115: 447-458.
28       Haynes, R. W.; Adams, D. M. (1992) Assessing economic impacts of air pollution damage to U.S. forests. In: de
29             Steiguer, J.  E., ed. The economic impact of air pollution on timber markets: studies from North America and
30             Europe. Asheville, NC: U.S. Department of Agriculture, Southeastern Forest Experiment Station; general
31             technical report no. SE-75.
32       Heagle,  A. S. (1979) Effects of growth media, fertiliser rate and hour and season of exposure on sensitivity of four
33             soybean cultivars to ozone. Environ. Pollut. 18: 313-322.
34       Heagle,  A. S.; Body, D. E.; Heck,  W. W. (1973) An open-top field chamber to assess the impact of air pollution on
35             plants. J. Environ. Qual. 2: 365-368.
36       Heagle,  A. S.; Body, D. E.; Neely,  G. E. (1974) Injury and yield responses of soybean to chronic doses of ozone and
37             sulfur dioxide in the field. Phytopathology 64: 132-136.
38       Heagle,  A. S.; Spencer, S.; Letchworth, M. B. (1979) Yield response of winter wheat to chronic doses of ozone. Can. J.
39             Bot. 57: 1999-2005.
40       Heagle,  A. S.; Letchworth, M. B.; Mitchell,  C. A.  (1983) Effects of growth medium and fertilizer rate on the yield
41             response of soybeans exposed to chronic doses of ozone. Phytopathology 73: 134-139.
42       Heagle,  A. S.; Heck, W. W.; Lesser, V. M.; Rawlings, J. O. (1987) Effects of daily ozone exposure duration and
43             concentration fluctuation on yield of tobacco. Phytopathology 77: 856-862.
44       Heagle,  A. S.; Kress, L. W.; Temple, P. J.; Kohut, R. J.; Miller, J. E.; Heggestad, H. E. (1988) Factors influencing
45             ozone dose-yield response relationships in open-top  field chamber studies. In: Heck, W. W.; Taylor, O. C.;
46             Tingey, D. T., eds. Assessment of crop loss from air pollutants: proceedings of an international conference;
47             October, 1987; Raleigh, NC. New York, NY: Elsevier Applied Science; pp.  141-179.
48       Heagle,  A. S.; Rebbeck, J.; Shafer, S. R.; Blum, U.; Heck, W. W. (1989) Effects of long-term ozone exposure and soil
49             moisture deficit on growth of a ladino clover-tall fescue pasture. Phytopathology 79:  128-136.
50       Heagle,  A. S.; McLaughlin, M. R.; Miller, J. E.; Joyner, R.  L.; Spruill, S. E.  (1991) Adaptation of a white clover
51             population to ozone stress. New Phytol. 119: 61-68.
52       Heagle,  A. S.; Miller, J. E.; Rawlings, J. O.; Vozzo, S. F. (1991) Effect of growth stage on soybean response to chronic
53             ozone exposure. J. Environ.  Qual.  20: 562-570.
         January 2005                                    9-363       DRAFT-DO NOT QUOTE OR CITE

-------
  1       Heagle, A. S.; Miller, J. E.; Sherrill. D. E.; Rawlings, J. O. (1993) Effects of ozone and carbon dioxide mixtures on
  2             two clones of white clover. New Phytol. 123: 751-762.
  3       Heagle, A. S.; Brandenburg, R. L.; Burns, J. C.; Miller, J. E. (1994) Ozone and carbon dioxide effects on spider mites
  4             in white clover and peanut. J. Environ. Qual. 23: 1168-1176.
  5       Heagle, A. S.; Miller, J. E.; Sherrill, D. E. (1994) A white clover system to estimate effects of tropospheric ozone on
  6             plants. J. Environ. Qual. 23: 613 -621.
  7       Heagle, A. S.; Miller, J. E.; Chevone, B. L; Dreschel, T. W.; Manning, W. J.; McCool, P. M.; Morrison, C. L.; Neely,
  8             G. E.; Rebbeck, J. (1995) Response of a white clover indicator system to tropospheric ozone at eight locations
  9             in the United States.  Water Air Soil Pollut. 85: 1373-1378.
10       Heagle, A. S.; Reinert, R. A.; Miller, J. E. (1996) Response of white clover to ozone in different environments. J.
11             Environ. Qual. 25: 273-278.
12       Heagle, A. S.; Miller, J. E.; Pursley, W. A. (1998) Influence of ozone stress on soybean response to carbon dioxide
13             enrichment: III. Yield and seed quality. Crop Sci. 38: 128-134.
14       Heagle, A. S.; Miller, J. E.; Booker, F. L.: Pursley, W. A. (1999) Ozone stress,  carbon dioxide enrichment, and
15             nitrogen fertility interactions in cotton. Crop Sci. 3 9: 731 -741.
16       Heagle, A. S.; Miller, J. E.; Pursley, W. A. (2000) Growth and yield responses of winter wheat to mixtures of ozone
17             and carbon dioxide. Crop Sci. 40:  1656-1664.
18       Heagle, A. S.; Stefanski, L. A. (2000) Relationships between ambient ozone regimes and white clover forage
19             production using different ozone exposure indexes. Atmos. Environ.  34: 735-744.
20       Heagle, A. S.; Miller, J. E.; Burkey, K. O.; Eason, G.; Pursley, W.  A. (2002) Growth and yield responses of snap bean
21             to mixtures of carbo  dioxide and ozone. J. Environ. Qual. 31: 2008-2014.
22       Heath, R. L. (1980) Initial events in injury to plants by air pollutants. Annu. Rev. Plant Physiol. 31: 395-431.
23       Heath, R. L. (1984) Air pollutant effects on biochemicals derived from metabolism: organic, fatty and amino acids. In:
24             Koziol, M. J.; Whatley, F. R., eds. Gaseous Air Pollutants and Plant Metabolism. London, United Kingdom:
25             Butterworths; pp. 275-290.
26       Heath, R. L. (1987) The biochemistry of ozone attack on the plasma membrane of plant cells. Recent Adv. Phytochem.
27             21:29-54.
28       Heath, R. L. (1988) Biochemical mechanisms of pollutant stress. In: Heck, W. W.; Taylor, O. C.; Tingey, D. T., eds.
29             Assessment of crop loss from air pollutants: proceedings of an international conference; October, 1987; Raleigh,
30             NC. London, United Kingdom: Elsevier Applied Science; pp. 259-286.
31       Heath, R. L. (1994a) Alterations of plant metabolism by ozone exposure. In: Alscher, R. G.; Wellburn, A. R., eds.
32             Plant Responses to the Gaseous Environment:  molecular, metabolic, and physiological aspects. London, United
33             Kingdom: Chapman & Hall; pp. 121-146.
34       Heath, R. L. (1994b) Possible mechanisms for the inhibition of photosynthesis by ozone. Photosynt. Res. 39: 439-451.
35       Heath, R. L. (1996) The modification of photo synthetic capacity induced by ozone exposure. In: Baker, N. R., ed.
36             Photosynthesis and the environment. Dordrecht, The Netherlands: Kluwer Academic Publishers; pp.  409-433.
37             (Advances in Photosynthesis, v.5).
38       Heath, R. L.; Castillo, F. J.  (1987) Membrane disturbances in response to  air pollutants. In: Schulte-Hostede, S.;
39             Darrall, N. M.; Blank, L. W.; Wellburn, A. R., eds. Air pollution and plant metabolism. New York, NY:
40             Elsevier Applied Science;  pp. 55-75.
41       Heath, R. L.; Taylor, G. E., Jr. (1997) Physiological processes and plant responses to ozone exposure. In: Sandermann,
42             H.; Wellburn, A. R.; Heath, R. L., eds. Forest Decline and Ozone: a comparison of controlled chamber and field
43             experiments. New York, NY: Springer-Verlag; pp. 317-368. (Ecological Studies, v. 127).
44       Heck, W. W.; Cowling, E. B. (1997) The need for a long term cumulative secondary ozone standard - an ecological
45             perspective. EM (January): 23-33.
46       Heck, W. W.; Dunning, J. A.; Hindawi, I. J.  (1966) Ozone: nonlinear relation of dose and injury in plants. Science
47             (Washington, DC) 151: 577-578.
48       Heck, W. W.; Philbeck, R. B.; Dunning, J. A. (1978) A continuous stirred tank reactor (CSTR) system for exposing
49             plants to gaseous air contaminants: principles,  specifications, construction, and operation. U.S. Agric. Res. Serv.
50             South. Reg. Rep. ARS-S-181.
51       Heck, W. W.; Taylor, O. C.; Adams, R.; Bingham, G.; Miller, J.; Preston, E.; Weinstein, L. (1982) Assessment of crop
52             loss from ozone. J. Air Pollut. Control Assoc. 32: 353-361.
         January 2005                                    9-364        DRAFT-DO NOT QUOTE OR CITE

-------
  1       Heck, W. W.; Heagle, A. S.; Miller, J. E.; Rawlings, J. O. (1991) A national program (NCLAN) to assess the impact of
  2             ozone on agricultural resources. In: Berglund, R. L.; Lawson, D. R.; McKee, D. J., eds. Tropospheric ozone and
  3             the environment: papers from an international conference; March 1990; Los Angeles, CA. Pittsburgh, PA: Air
  4             & Waste Management Association; pp. 225-254. (A&WMA transactions series no. TR-19).
  5       Heck, W. W.; Furiness, C. S.; Cowling, E. B.; Sims, C. K. (1998) Effects of ozone on crop, forest, and natural
  6             ecosystems: assessment of research needs. EM (October): 11-22.
  7       Heggestad, H. E. (1988) Reduction in soybean seed yields by ozone air pollution? JAPCA 38: 1040-1041.
  8       Heliovaara, K.; Vaisanen, R. (1993) Insects and pollution. Boca Raton, FL: CRC Press.
  9       Hendrey, G. R.; Kimball, B. A. (1994) The FACE program. Agric. For. Meteorol. 70: 3-14.
10       Hendrey, G. R.; Ellsworth, D. S.; Lewin, K. F.; Nagy, J. (1999) A free-air enrichment system for exposing tall forest
11             vegetation to elevated atmospheric CO2.  Global Change Biol. 5: 293-309.
12       Hertstein, U.; Fangmeier, A.; Jager, H. J. (1996) ESPACE-wheat (European stress physiology and climate experiment
13             - Project  1: wheat): objectives, general approach, and first results. J. Appl. Bot. 70: 172-180.
14       Hertstein, U.; Colls, J.; Ewert, F.; Van Oijen, M. (1999) Climatic conditions and concentrations of carbon dioxide and
15             air pollutants during 'ESPACE-wheat' experiments. Eur. J. Agron.10: 163-169.
16       Hildebrand, E.; Skelly, J. M.; Fredericksen, T. S. (1996) Foliar response of ozone-sensitive hardwood tree species from
17             1991 to 1993 in the Shenandoah National Park, Virginia. Can. J. For.  Res. 26: 658-669.
18       Hogsett, W. E.; Plocher, M.; Wildman, V.; Tingey, D. T.; Bennett, J. P. (1985) Growth response of two varieties of
19             slash pine seedlings to chronic ozone exposures. Can. J. Bot. 63: 2369-2376.
20       Hogsett, W. E.; Tingey, D. T.; Herman, S. R. (1985) A programmable exposure control system for determination of the
21             effects of pollutant exposure regimes on plant growth. Atmos. Environ. 19: 1135-1145.
22       Hogsett, W. E.; Olszyk, D.; Ormrod, D. P.; Taylor, G. E., Jr.; Tingey, D. T. (1987a) Air pollution exposure systems
23             and experimental protocols: volume 1: a review and evaluation of performance. Corvallis, OR: U.S.
24             Environmental Protection Agency, Environmental Research Laboratory; EPA report no. EPA/600/3-87/037a.
25             Available from:  NTIS, Springfield, VA; PB88-181680.
26       Hogsett, W. E.; Olszyk, D.; Ormrod, D. P.; Taylor, G. E., Jr.; Tingey, D. T. (1987b) Air pollution exposure systems
27             and experimental protocols: volume 2: description of facilities. Corvallis, OR: U.S. Environmental Protection
28             Agency, Environmental Research Laboratory; EPA report no. EPA/600/3-87/037b. Available from: NTIS,
29             Springfield, VA; PB88-181698.
30       Hogsett, W. E.; Tingey, D. T.; Lee, E. H. (1988) Ozone exposure indices:  concepts for development and evaluation of
31             their use. In: Heck, W. W.; Taylor, O. C.; Tingey,  D. T., eds.  Assessment of crop loss from air pollutants:
32             proceedings of an international conference; October 1987; Raleigh, NC. New York, NY: Elsevier Applied
33             Science; pp. 107-138.
34       Hogsett, W. E.; Herstrom, A. A.; Laurence, J. A.; Lee, E. H.; Weber, J. E.; Tingey, D. T.  (1995) Risk characterization
35             of tropospheric ozone to forests. In: Lee,  S. D.; Schneider, T., eds. Comparative risk analysis and priority setting
36             for air pollution issues: proceedings of the 4th U.S.-Dutch international symposium; June 1993; Keystone, CO.
37             Pittsburgh, PA: Air & Waste Management Association; pp. 119-145.  (AWMA publication no. VTP-43).
38       Hogsett, W. E.; Weber, J. E.; Tingey, D.; Herstrom, A.; Lee, E. H.; Laurence, J. A. (1996) An approach for
39             characterizing tropospheric ozone  risk to  forests. Environ. Manage. (N. Y.) 21: 105-120.
40       Hogsett, W. E.; Weber, J. E.; Tingey, D.; Herstrom, A.; Lee, E. H.; Laurence, J. A. (1997) Environmental auditing: an
41             approach for characterizing tropospheric  ozone risk to forests. Environ. Manage. 21: 105-120.
42       Holland, D. M.; Saltzman, N.; Cox, L. H.; Nychka, D. (1999) Spatial prediction of sulfur dioxide in the eastern United
43             States. In: Gomez-Hernandez, J.; Scares,  A.; Froidevaux, R., eds. geoENV II: geostatistics for environmental
44             applications, proceedings of the second European conference; November 1998; Valencia, Spain. Dordrecht, The
45             Netherlands: Kluwer Academic Publishers; pp. 65-76. (Quantitative geology and geostatistics: v. 10).
46       Holopainen, J. K.; Kainulainen, P.; Oksanen, J.  (1997) Growth and reproduction of aphids and levels of free amino
47             acids in Scots pine and Norway  spruce in an open-air fumigation with ozone. Global Change Biol. 3:  139-147.
48       Holton, M. K.; Lindroth, R. L.; Nordheim, E. V. (2003) Foliar quality influences tree-herbivore-parisitoid interactions:
49             effects of elevated CO2, O3, and plant genotype. Oecologia 137: 233-244.
50       Horemans, N.; Foyer, C. H.; Asard, H. (2000) Transport and action of ascorbate at the plant plasma membrane.  Trends
51             Plant Sci. 5:263-267.
52       Hormaza, J. I.; Pinney, K.; Polito, V. S. (1996) Correlation in the tolerance to ozone between sporophytes and male
53             gametophytes of several fruit and tree nut species (Rosaceae). Sex.  Plant Reprod. 9: 44-48.
         January 2005                                    9-365        DRAFT-DO NOT QUOTE OR CITE

-------
  1       Houpis, J. L. I; Costella, M. P.; Cowles, S. (1991) A branch exposure chamber for fumigating ponderosa pine to
  2             atmospheric pollution. J. Environ. Qual. 20: 467-474.
  3       Hui, D.; Sims, D. A.; Johnson, D. W.; Chang, W.; Luo, Y. (2002) Effects of gradual versus step increases in carbon
  4             dioxide onPlantago photosynthesis and growth in a microcosm study. Environ. Exp. Bot. 47: 51-66.
  5       Hummel, R. L.; Brandenburg, R. L.; Heagle, A. S.; Arellano, C. (1998) Effects of ozone on reproduction of
  6             two-spotted spider mite (Acari: Tetrannychidae)  on white clover. Environ. Entomol. 27: 388-394.
  7       Idso, K. E.; Idso, S. B. (1994) Plant responses to atmospheric CO2 enrichment in the face of environmental constraints:
  8             a review of the past 10 years' research. Agric. For. Meteorol. 69: 153-203.
  9       Innes, J. L.; Skelly, J. M.; Schaub, M. (2001) Ozone and broadleaved species. A guide to the identification of
10             ozone-induced foliar injury.  Bern, Switzerland: Paul Haupt Publishers.
11       Intergovernmental Panel on Climate Change (IPCC). (1996) Climate change 1995: the science of climate change:
12             contribution of working group I to the second assessment of the Intergovernmental Panel on Climate Change.
13             Cambridge, United Kingdom: Cambridge University Press.
14       Intergovernmental Panel on Climate Change (IPCC). (1998) The regional impacts of climate change: an assessment of
15             vulnerability, a special report of IPCC Working Group II. Watson, R. T.; Zinyowera, M. C.; Moss, R. H., eds.
16             Cambridge, United Kingdom: Cambridge University Press.
17       Isebrands, J. G.; Dickson, R. E.; Rebbeck, J.; Karnosky, D. F. (2000) Interacting effects of multiple stresses on growth
18             and physiological processes  in northern forest trees. In: Mickler, R. A.; Birsdey, R. A.; Horn, J., eds. Responses
19             of northern U.S. forests to environmental change. New York, NY: Springer-Verlag; pp. 149-180. (Ecological
20             studies: v. 139).
21       Isebrands, J. G.; McDonald, E. P.; Kruger, E.; Hendrey, G.; Percy, K.; Pregitzer, K.; Sober, J.; Karnosky, D. F. (2001)
22             Growth responses ofPopulus tremuloides clones to interacting carbon dioxide and tropospheric ozone. Environ.
23             Pollut.  115:359-371.
24       Izuta, T. (1998) Ecopysiological responses of Japanese  forest tree species to ozone, simulated acid rain and soil
25             acidification.  J. Plant Res. Ill: 471-480.
26       Jackson, G. E. (1995) The effect of ozone, nitrogen dioxide and nitric oxide fumigation of cereals on the rose grain
27             aphid Metopolophium dirhodum. Agric. Ecosyst. Environ.  54: 187-194.
28       Jackson, D. M.; Heagle, A.  S.; Eckel, R. V. W. (1999) Ovipositional response of tobacco hornworm moths
29             (Lepidoptera: Sphyngidae) to tobacco plants grown under elevated levels  of ozone. Environ. Entomol. 28:
30             566-571.
31       Jackson, D. M.; Rufty, T. W.; Heagle, A. S.; Severson,  R. F.; Eckel, R. W. V. (2000) Survival and development of
32             tobacco hornworm larvae on tobacco plants grown under elevated levels of ozone. J. Chem. Ecol. 26: 1-18.
3 3       Jacobson, J. S. (1977) The effects of photochemical oxidants on vegetation. In: Ozon und Begleitsubstanzen im
34             photochemischen Smog: das Kolloquium  [Ozone and related substances in photochemical smog: the
35             colloquium];  September 1976; Dtisseldorf, Federal Republic of Germany. Dtisseldorf, Federal Republic of
36             Germany: VDI-Verlag GmbH; pp.  163-173. (VDI-Berichte nr. 270).
37       Jager, H. J.; Herstein, U.; Fangmeier, A. (1999) The European Stress Physiology and Climate Experiment - Project 1:
38             wheat (ESPACE-wheat); introduction, aims and  methodology. Eur. J. Agron. 10: 155-162.
39       James, R. L.; Cobb, F. W., Jr.; Miller, P. R.; Parmeter, J. R., Jr. (1980) Effects of oxidant air pollution on susceptibility
40             of pine roots to Fames annosus. Phytopathology  70: 560-563.
41       Janson, R.; Granat, L. (1999) A foliar rinse study of the dry deposition of nitric acid to a coniferous forest. Agric. For.
42             Meteorol. 98-99: 683-696.
43       Jarvis, P. G. (1976) The interpretation of the variations  in leaf water potential and stomatal conductance found in
44             canopies in the field. Philos. Trans. R. Soc. B 273: 593-610.
45       Jiang, C.-Z.; Rodermel, S. R.; Shibles, R. M. (1993) Photosynthesis, rubisco activity and amount, and their regulation
46             by transcription in senescing soybean leaves. Plant Physiol. 101: 105-112.
47       Jimenez, A.; Calvo, E.; Martin, C.;  Porcuna, J. L.; Sanz, M. J. (2001) Estudio de la interaccion entre el ozone y el
48             desarrolio de las virosis en el cultivo del tomate.  Agricola Vergel 231: 141-150.
49       Johnson, B. G.; Hale, B. A.; Ormrod, D. P. (1996a) Carbon dioxide and ozone effects on growth of a legume-grass
50             mixture. J. Environ. Qual.  25: 908-916.
51       Johnson, J. D.; Chappelka, A. H.; Hain, F. P.; Heagle, A. S. (1996b) Interactive effects of air pollutants with abiotic
52             and biotic factors on southern pine forests. In:  Cox, S.; Mickler, R. A., eds. Impact of air pollutants on southern
53             pine forests. New York, NY: Springer; pp. 281-312. (Ecological studies: v. 118).
         January 2005                                    9-366        DRAFT-DO NOT QUOTE OR CITE

-------
  1       Jones, H. G.; Lacasse, N. L.; Liggett, W. S.; Weatherford, F. (1977) Experimental air exclusion system for field studies
  2             of SO2 effects on crop productivity. Washington, DC: U.S. Environmental Protection Agency; report no.
  3             EPA-600/7-77-122.
  4       Jones, C. G.; Coleman, J. S.; Findlay, S. (1994) Effects of ozone on interactions between plants, consumers and
  5             decomposers. In: Alscher, R. G.; Wellburn, A. R., eds. Plant responses to the gaseous environment. Molecular,
  6             metabolic and physiological aspects. London, United Kingdom: Chapman and Hall; pp. 339-363.
  7       Joss, U.; Graber, W. K. (1996) Profiles and simulated exchange of H2O, O3, NO2 between the atmosphere and the
  8             HartX Scots pine plantation. Theor. Appl. Climatol. 53: 157-172.
  9       Kainulainen, P.; Holopainen, J. K.; Hyttinen, H.; Oksanen, J. (1994) Effect of ozone on the biochemistry and aphid
10             infestation of Scots pine. Phytochemistry 35: 39-42.
11       Kainulainen, P.; Holopainen, J. K.; Holopainen, T. (2000a) Combined effects of ozone and nitrogen on secondary
12             compounds, amino acids, and aphid performance in Scots pine. J. Environ. Qual. 29: 334-342.
13       Kainulainen, P.; Utriainen, J.; Holopainen, J. K.; Oksanen, J.; Holopainen, T. (2000b) Influence of elevated ozone and
14             limited nitrogen availability on conifer seedlings in an open-air fumigation system: effects on growth, nutrient
15             content, mycorhiza, needle ultrastructure, starch and secondary compounds. Global Change Biol. 6: 345-355.
16       Kainulainen, P.; Holopainen, T.; Holopainen, J. K. (2003) Decomposition of secondary compounds from needle litter
17             of Scots pine grown under elevated CO2 and O3. Glob. Change Biol. 9: 295-304.
18       Kangasjarvi, J.; Talvinen, J.; Utriainen, M; Karjalainen, R. (1994) Plant defense systems induced by ozone. Plant Cell
19             Environ. 17: 783-794.
20       Karenlampi, L.; Skarby, L. (1996) Critical levels for ozone in Europe: testing and finalizing the concepts. UN-ECE
21             workshop report. In: Proceedings of UN-ECE convention on long-range transboundary air pollution workshop;
22             April; Kuopio, Finland. Kupio, Finland: University of Kuopio, Department of Ecology and Environmental
23             Science.
24       Karlsson, P. E.; Medin, E.-L.; Wickstrom, H.; Sellden, G.; Wallin, G.; Ottosson, S.; Skarby, L. (1995) Ozone and
25             drought stress: interactive effects on the growth and physiology of Norway spruce (Picea abies (L.) Karst).
26             Water Air Soil Pollut. 85:  1325-1330.
27       Karlsson, P. E.; Medin, E. L.; Wallin, G.; Sellden, G.;  Skarby, L. (1997) Effects of ozone and drought stress on the
28             physiology  and growth of two clones of Norway spruce (Picea abies). New Phytol. 136: 265-275.
29       Karlsson, P. E.; Plaijel, H.; Karlsson, G. P.; Medin, E.  L.; Skarby, L. (2000) Simulations of stomatal conductance and
30             ozone uptake to Norway spruce saplings in open-top chambers. Environ. Pollut. 109:  443-451.
31       Karlsson, P. E.; Medin, E. L.; Sellden, G.; Wallin, G.;  Ottosson,  S.; Pleijel, H.; Skarby, L. (2002) Impact of ozone and
32             reduced water supply on the biomass accumulation of Norway spruce saplings. Environ. Pollut. 19: 237-244.
33       Karlsson, P. E.; Sellden, G.; Plaijel, H. (2003) Establishing ozone critical levels II. UNECE workshop report.
34             Gothenburg, Sweden: IVL Swedish Environmental Research Institute; IVL report B 1523 A,B,C. Available:
35             http://www.ozoneworkshop.ivl.se/wsreport.htm [7 July, 2004].
36       Karlsson, G. P.; Karlsson, P. E.; Danielsson, H.; Pleijel, H. (2003) Clover as a tool for bioindication of phytotoxic
37             ozone-5 years of experience from southern Sweden-consequences for the short-term critical levels. Sci. Total
38             Environ. 301:205-213.
39       Karlsson, P. E.; Uddling, J.; Skarby,  L.; Wallin, G.; Sellden, G. (2003) Impact of ozone on the growth of birch (Betula
40             pendula) saplings. Environ. Pollut. 124: 485-495.
41       Karnosky, D. F. (1981) Changes in eastern white pine  stands related to air pollution stress. Mitt. Forstl.
42             Bundesversuchsanst. Wien 137: 41-45.
43       Karnosky, D. F. (2003) Impacts of elevated atmospheric CO2 on forest trees and forest ecosystems: knowledge gaps.
44             Environ. Int. 29: 161-169.
45       Karnosky, D. F.; Gagnon, Z. E.; Dickson, R. E.; Coleman, M. D.; Lee, E. H.; Isebrands, J. G. (1996) Changes in
46             growth, leaf abscission, biomass associated with seasonal tropospheric ozone exposures ofPopulus tremuloides
47             clones and seedlings. Can. J. For. Res. 26: 23-37.
48       Karnosky, D. F.; Podila, G. K.; Gagnon, Z.; Pechter, P.; Akkapeddi, A.; Coleman, M.; Dickson, R. E.; Isebrands, J. G.
49             (1998) Genetic control of responses to interacting O3 and CO2 in Populus tremuloides. Chemosphere 36:
50             807-812.
51       Karnosky, D. F.; Mankovska, B.; Percy, K.; Dickson, R. E.; Podila, G. K.; Sober, J.; Noormets, A.; Hendrey, G.;
52             Coleman, M. D.; Kubiske, M.; Pregitzer, K. S.;  Isebrands, J. G. (1999) Effects of tropospheric ozone on
         January 2005                                    9-367       DRAFT-DO NOT QUOTE OR CITE

-------
  1             trembling aspen and interaction with CO2: results from an O3-gradient and a FACE experiment. Water Air Soil
  2             PoUut. 116:311-322.
  3       Karnosky, D. F.; Ceulemans, R.; Scarascia-Mugnozza, G. E.; Ines, J. L., eds. (2001) The impact of carbon dioxide and
  4             other greenhouse gases on forest ecosystems: report no. 3 of the IUFRO Task Force on Environmental Change.
  5             Wallingford, United Kingdom: CABI Publishing. (IUFRO research series no. 8). Available:
  6             http://www.cabi-publishing.org/Bookshop/Readingroom/0851995519.asp [19 August, 2003].
  7       Karnosky, D. F.; Gielen, B.; Ceulemans, R.; Schlesinger, W. H.; Norby, R. J.; Oksanen, E.; Matyssek, R.; Hendrey, G.
  8             R. (200 la) FACE systems for studying the impacts of greenhouse gases on forest ecosystems. In: Karnosky, D.
  9             F.; Ceulemans, R.; Scarascia-Mugnozza, G. E.; Innes, J. L. eds. The impact of carbon dioxide and other
10             greenhouse gases on forest ecosystems; report no. 3 of the IUFRO Task Force on Environmental Change.
11             Wallingford, UK: CABI Publishers, pp. 297-324. Available:
12             http://www.cabi-publishing.org/Bookshop/Readingroom/0851995519.asp [16 April, 2004].
13       Karnosky, D. F.; Oksanen, E.; Dickson, R. E.; Isebrands, J. G. (200 Ib) Impact of interacting greenhouse gases on
14             forest ecosystems. In: Karnosky, D. F.; Ceulemans, R.; Scarascia-Mugnozza, G. E.; Innes, J. L., eds. The impact
15             of carbon dioxide and other greenhouse gases of forest ecosystems:  report no. 3 of the IUFRO Task Force on
16             Environmental Change. Wallingford, United Kingdom: CABI Publishing; pp. 253-267. (IUFRO research series
17             no. 8).
18       Karnosky, D. F.; Percy, K. E.; Xiang, B.; Callan, B.; Noormets, A.; Mankovska, B.; Hopkin, A.; Sober, J.; Jones, W.;
19             Dickson, R. E.; Isebrands, J. G. (2002) Interacting elevated CO2 and tropospheric O3 predisposes aspen
20             (Populus tremuloides Michx.) to infection by rust (Melampsora medusae f. sp. tremuloidae). Global Change
21             Biol. 8: 329-338.
22       Karnosky, D. F.; Zak, D. R.; Pregitzer, K. S.; Awmack, C.  S.; Bockheim, J. G.; Dickson, R. E.; Hendrey, G. R.; Host,
23             G. E.; King, J. S.; Kopper, B. J.; Kruger, E. L.; Kubiske, M. E.; Lindroth, R. L.; Mattson, W. J.; McDonald, E.
24             P.; Noormets, A.; Oksanen, E.; Parsons, W. F. J.; Percy, K. E.; Podila, G. K.; Riemenschneider, D. E.; Sharma,
25             P.; Thakur, R.; Sober, A.; Sober, J.; Jones, W. S.; Anttonen, S.; Vapaavuori, E.; Mankovska, B.; Heilman, W.;
26             Isebrands, J. G. (2003a) Tropospheric O3 moderates responses of temperate hardwood forests to elevated CO2:
27             A synthesis of molecular to ecosystem results from the Aspen FACE project. Funct. Ecol. 17: 289-304.
28       Karnosky, D. F.; Percy, K.; Mankovska, B.; Pritchard, T.; Noormets, A.; Dickson, R. E.; Jepsen, E.; Isebrands, J. G.
29             (2003b)  Ozone affects the fitness of trembling aspen. In: Karnosky,  D. F.; Percy, K. E.; Chappelka, A. H.;
30             Simpson, C. J.; Pikkarainen, J., eds. Air pollution, global change and forests in the new millenium. Oxford,
31             United Kingdom: Elsevier Science. (Developments in Environmental Science, 3).
32       Karpinski, S.; Wingsle, G.; Karpinska, B.; Hallgren, J.-E. (1993) Molecular response to photooxidative stress in Firms
33             sylvestris (L.): II. Differential expression of Cu-Zn superoxide dismutases and glutathione reductase. Plant
34             Physiol.  103: 1385-1391.
3 5       Kasurinen, A. E.; Helmisaari, H.-S.; Holopainen, T. (1999) The influence of elevated CO2 and O3 on fine roots and
3 6             mycorrihizas of naturally growing young Scots pine trees during three exposure years. Global Change Biol. 5:
37             771-780.
38       Kats, G.; Thompson, C. R.; Kuby,  W. C.  (1976) Improved ventilation of open top greenhouses. J. AirPollut. Control
39             Assoc. 26: 1089-1090.
40       Kats, G.; Dawson, P. J.; Bytnerowicz, A.; Wolf,  J. W.; Thompson, C. R.; Olszyk, D. M. (1985) Effects of ozone or
41             sulfur dioxide on growth and yield of rice. Agric. Ecosyst. Environ.  14:  103-117.
42       Kellomaki, S.; Wang, K.-Y. (1997a) Effects of elevated O3 and CO2 concentrations on photosynthesis and stomatal
43             conductance in Scots pine. Plant Cell Environ. 20: 995-1006.
44       Kellomaki, S.; Wang, K.-Y. (1997b) Effects of elevated O3 and CO2 on chlorophyll fluorescence and gas exchange in
45             Scots pine during the third growing season. Environ. Pollut. 97: 17-27.
46       Kelting, D. L.; Burger, J. A.; Edwards, G. S. (1995) The effects of ozone on the root dynamics of seedlings and mature
47             red oak (Quercus rubra L.). For. Ecol. Manage. 79:  197-206.
48       Kerstiens, G.; Lendzian, K. J. (1989) Interaction between ozone and plant cuticles: I. ozone deposition and
49             permeability. New Phytol. 112: 13-19.
50       Khan, M. R.; Khan, M. W. (1994)  Single and interactive effects of O3 and  SO2 on tomato. Environ. Exp. Bot. 34:
51             461-469.
52       Khan, M. R.; Khan, M. W. (1997)  Effects of the root-knot nematode, Meloidogyne incognita, on the sensitivity of
53             tomato to sulfur dioxide and ozone. Environ. Exp. Bot. 38: 117-130.
         January 2005                                   9-368        DRAFT-DO NOT QUOTE OR CITE

-------
  1       Khan, M. R.; Khan, M. W. (1998a) Interactive effects of ozone and powdery mildew (Sphaerotheca fuliginea) on
  2             bottle gourd (Lagenaria siceraria). Agric. Ecosyst. Environ. 70: 109-118.
  3       Khan, M. R.; Khan, M. W. (1998b) Interactive effects of ozone and root-knot nematode on tomato. Agric. Ecosyst.
  4             Environ. 70: 97-107.
  5       Khan, M. R.; Khan, M. W. (1999) Effects of intermittent ozone exposures on powdery mildew of cucumber. Environ.
  6             Exp. Bot. 42:  163-171.
  7       Kickert, R. N.; Krupa, S. V. (1991) Modeling plant response to tropospheric ozone: a critical review. Environ. Pollut.
  8             70:271-383.
  9       Kickert, R. N.; Tonella, G.; Simonov, A.; Krupa, S. V. (1999) Predictive modeling of effects under global change.
10             Environ. Pollut. 100: 87-132.
11       Kieliszewska-Rokicka, B.; Rudawska, M.; Leski, T. (1998) Ectomycorrhizae of young and mature Scots pine trees in
12             industrial regions in Poland. In: Bytnerowicz, A.; Arbaugh, M. J.; Schilling, S. L., eds. Proceedings of the
13             International Symposium on Air Pollution and Climate Change Effects on Forest Ecosystems; February, 1996;
14             Riverside, CA. Albany, CA: USDA Forest Service, Pacific Southwest General Technical Report no. PSW-GTR
15             166; pp. 205-214.
16       Kim, H. J.; Helfand,  G. E.; Howitt, R. E. (1998) An economic analysis of ozone control in California's San Joaquin
17             Valley. J. Agric. Resource Econ.  23: 55-70.
18       Kim, J. S.; Chappelka, A. H.; Miller-Goodman, M. S. (1998) Decomposition of blackberry and broomsedge bluestem
19             as influenced by ozone. J. Environ. Qual. 27: 953-960.
20       King, J. S.; Pregitzer, K. S.; Zak, D. R.;  Sober, J.; Isebrands, J. G.; Dickson, R. E.; Hendrey, G. R.; Karnosky,  D. F.
21             (2001) Fine-root biomass and fluxes of soil carbon in young stands of paper birch and trembling aspen as
22             affected by elevated atmospheric CO2 and tropospheric O3. Oecologia 128: 237-250.
23       Kleier, C.; Farnsworth, B.; Winner, W. (1998) Biomass, reproductive output, and physiological responses of
24             rapid-cycling Brassica (Brassica rapd) to ozone and modified root temperature. New Phytol.  139: 657-664.
25       Kleier, C.; Farnsworth, B.; Winner, W. (2001) Photosynthesis and biomass allocation of radish cv. "Cherry Belle" in
26             response to root temperature and  ozone. Environ. Pollut. Ill: 127-133.
27       Kliebenstein, D. J.; Monde, R.-A.; Last, R. L. (1998) Superoxide dismutase in Arabidopsis: an eclectic  enzyme family
28             with disparate regulation and protein localization. Plant Physiol. 118: 637-650.
29       Kobayashi, K.; Okada, M.; Nouchi, I. (1994) A chamber system for exposing rice (Oryza sativa L.) to ozone in a
30             paddy field. New Phytol. 126: 317-325.
31       Kobayashi, K.; Okada, M.; Nouchi, I. (1995) Effects of ozone on dry matter partitioning and yield of Japanese
32             cultivars of rice (Oryza sativa L.). Agric. Ecosyst. Environ. 53: 109-122.
33       Kohut, R. J.; Laurence, J. A.; Amundson, R. G. (1988) Effects of ozone and sulfur dioxide on yield of red clover and
34             timothy. J. Environ. Qual. 17: 580-585.
35       Kohut, R. J.; Laurence, J. A.; King, P.; Raba, R. (1997) Identification of bioindicator species for ozone  and assessment
36             of the responses to ozone of native vegetation at Acadia National Park. Denver CO: National Park Service; Air
37             Resources Division; NFS technical report Acadia D-175.
38       Kolb, T. E.; Matyssek, R. (2001) Limitations and perspectives  about scaling ozone impacts in trees. Environ. Pollut.
39             115:373-393.
40       Kolb, T. E.; Fredericksen, T. S.; Steiner, K. C.; Skelly, J. M. (1997) Issues in scaling tree size and age responses to
41             ozone: a review. Environ. Pollut.  98: 195-208.
42       Kollner, B.; Krause, G. H. M. (2003) Effects of two different ozone exposure regimes on chlorophyll and sucrose
43             content of leaves and yield parameters of sugar beet (Beta Vulgaris L.) and rape (Brassica Napus L.). Water Air
44             Soil Pollut. 144: 317-332.
45       Kolomiets, M. V.; Chen, H.; Gladon, R. J.; Braun, E. J.; Hannapel, D. J. (2000) A leaf lipoxygenase of potato induced
46             specifically by pathogen infection. Plant Physiol. 124: 1121-1130.
47       Kopp, R. J.; Vaughn, W. J.; Hazilla, M.; Carson, R. (1985) Implications of environmental policy for U.S. agriculture:
48             the case of ambient ozone standards. J. Environ. Manage. 20: 321-331.
49       Kopper, B. J.; Lindroth, R. L. (2003) Responses of trembling aspen (Populus tremuloides) phytochemistry and aspen
50             blotch leafminer (Phyllonorycter  tremuloidiella) performance to elevated levels of atmospheric CO2 and O3.
51             Agric. For. Entomol. 5: 17-26.
52       Kopper, B. J.; Lindroth, R. L. (2003) Effects of elevated carbon dioxide and ozone on the phytochemistry of aspen and
53             performance of an herbivore. Oecologia 134: 95-103.
         January 2005                                    9-369       DRAFT-DO NOT QUOTE OR CITE

-------
  1       Kostka-Rick, R.; Manning, W. J. (1992) Partitioning of biomass and carbohydrates in field-grown radish under
  2             ambient concentrations of ozone, and treated with the anti-ozonant ethylene-diurea (EDU). New Phytol. 121:
  3             187-200.
  4       Kostka-Rick, R.; Manning, W. J. (1993) Dose-response studies with the antiozonant ethylenediurea (EDU), applied as
  5             a soil drench to two growth substrates, on greenhouse-grown varieties ofPhaseolus vulgaris L. Environ. Pollut.
  6             82:63-72.
  7       Kostka-Rick, R.; Manning, W. J. (1993) Dynamics of growth and biomass partitioning in field-grown bush bean
  8             (Phaseolus vulgaris L.), treated with the antiozonant ethylenediurea (EDU). Agric. Ecosyst. Environ. 47:
  9             195-214.
10       Kostka-Rick, R.; Manning, W. J.; Buonaccorsi, J. P. (1993) Dynamics of biomass partitioning in field-grown radish
11             varieties, treated with ethylenediurea. Environ. Pollut. 80: 133-145.
12       Kouterick, K. B.; Skelly, J. M.; Fredericksen, T. S.; Steiner, K. C.; Kolb, T. E.; Ferdinand, J. A. (2000) Foliar injury,
13             leaf gas exchange and biomass responses of black cherry (Prunus serotina Ehrh.) half-sibling families to ozone
14             exposure. Environ. Pollut. 107:  117-126.
15       Koziol, M. J.; Whatley, F. R. (1984) Gaseous air pollutants and plant metabolism. London, United Kingdom:
16             Butterworths Scientific.
17       Kozlowski, T. T.; Kramer, P. J.; Pallardy, S. G. (1991) The physiological ecology of woody plants. San Diego, CA:
18             Academic Press, Inc.; pp. 338-375. (Mooney, H. A., ed. Physiological ecology: a series of monographs, texts,
19             and treatises).
20       Kraft, M.; Weigel, H.-J.; Mejer, G.-J.; Brandes, F. (1996) Reflectance measurements of leaves for detecting visible and
21             non-visible ozone damage to crops. J. Plant Physiol. 148: 148-154.
22       Krapp, A.; Hofmann, B.; Schafer, C.; Stitt, M. (1993) Regulation of the  expression of rbcS and other photosynthetic
23             genes by carbohydrates: a mechanism for the "sink regulation" of photosynthesis? Plant J. 3: 817-828.
24       Kress, L. W.; Miller,  J. E.; Smith, H. J. (1985) Impact of ozone on winter wheat yield. Environ. Exp.  Bot. 25: 211-228.
25       Krupa, S. V.; Nosal, M.; Legge, A. H.  (1994) Ambient ozone and crop loss: establishing a cause-effect relationship.
26             Environ. Pollut. 83:269-276.
27       Krupa, S. V.; Griinhage, L.; Jager, H.-J.; Nosal, M.; Manning, W. J.; Legge, A. H.; Hanewald, K. (1995) Ambient
28             ozone (O3) and adverse crop response: a unified view of cause and effect. Environ. Pollut. 87:  119-126.
29       Krupa, S. V.; Jager, H.-J. (1996) Adverse effects of elevated levels of ultraviolet (UV)-B radiation and ozone (O3) on
30             crop growth and productivity. In: Bazzaz, F.; Sombroek, W., eds. Global climate change and agricultural
31             production. Chichester, United Kingdom: John Wiley & Sons; pp. 141-169. Available:
32             http://www.fao.org/docrepAV5183E/w5183eOO.htm [28 August, 2003].
33       Krupa, S. V.; Kickert, R. N.; Jager, H.-J. (1998) Elevated ultraviolet (UV)-B radiation and agriculture. New York, NY:
34             Springer-Verlag Environmental Intelligence Unit.
3 5       Krupa, S. V.; Groth, J. V. (2000) Global climate change and crop responses: uncertainties associated with the current
36             methodologies. In: Agrawal, S.  B.; Agrawal, M., eds. Environmental pollution and plant responses. Boca Raton,
37             FL: Lewis Publishers; pp. 1-18.
38       Krupa, S. V.; Legge, A. H. (2000) Passive sampling of ambient, gaseous air pollutants: an assessment from an
39             ecological perspective. Environ. Pollut. 107:  31-45.
40       Krupa, S. V.; Nosal, M.; Peterson, D. L. (2001) Use of passive ozone O3 samplers in vegetation effects assessment.
41             Environ. Pollut. 112: 303-309.
42       Krupa, S. V.; Nosal, M.; Ferdinand, J.  A.; Stevenson, R. E.; Skelly, J. M. (2003) A multi-variate  statistical model
43             integrating passive sampler and meteorology data to predict the frequency distributions of hourly ambient ozone
44             (O3) concentrations. Environ. Pollut. 124: 173-178.
45       Kuehler, E. A.; Flagler, R. B. (1999) The effects of sodium erythorbate and ethylenediurea on photosynthetic function
46             of ozone-exposed loblolly pine  seedlings. Environ. Pollut. 105: 25-35.
47       Kuik, O. J.; Helming, J. F. M.; Dorland, C.; Spaninks, F. A. (2000) The  economic benefits to agriculture of a reduction
48             of low-level ozone pollution in The Netherlands. Eur. Rev. Agric Econ. 27: 75-90.
49       Kull, O.; Tulva, I.; Vapaavuori, E. (2003) Influence of elevated CO2 and O3 on Betula pendula Roth crown structure.
50             Ann. Bot. 91:559-569.
51       Kurczynska, E. U.; Dmuchowski, W.;  Wloch, W.; Bytnerowicz. A. (1997) The influence of air pollutants on needles
52             and stems of scots pine (Pinus Sylvestris L.) trees. Environ.  Pollut. 98: 325-334.
53       Kurpius, M. R.; McKay, J.; Goldstein, A. H. (2002) Annual ozone deposition to a Sierra Nevada ponderosa pine
54             plantation. Atmos. Environ. 36: 4503-4515.


         January 2005                                    9-370        DRAFT-DO NOT QUOTE OR CITE

-------
  1       Kytoviita, M.-M; Pelloux, J.; Fontaine, V.; Botton, B.; Dizengremel, P. (1999) Elevated CO2 does not ameliorate
  2             effects of ozone on carbon allocation in Pinus halapensis and Betula pendula in symbiosis with Paxillus
  3             involutes. Physiol. Plant. 106: 370-377.
  4       Kytoviita, M.-M.; Le Thiec, D.; Dizengremel, P. (2001) Elevated CO2 and ozone reduce nitrogen acquisition by Pinus
  5             halepensis from its mycorrhizal symbiont. Physiol. Plant. Ill: 305-312.
  6       Lagrimini, L. M.; Rothstein, S. (1987) Tissue specificity of tobacco peroxidase isozymes and their induction by
  7             wounding and tobacco mosaic virus infection. Plant Physiol. 84: 438-442.
  8       Laisk, A.; Kull, O.; Moldau, H. (1989) Ozone concentration in leaf intercellular air spaces is close to zero. Plant
  9             Physiol. 90: 1163-1167.
10       Lamb, C.; Dixon, R. A. (1997) The oxidative burst in plant disease resistance. Ann. Rev. Plant Physiol. Mol. Biol. 48:
11             251-275.
12       Landolt, W.; Gunthardt-Goerg, M. S.; Pfenninger, L; Einig, W.; Hampp, R.; Maurer, S.; Matyssek, R. (1997) Effect of
13             fertilization on ozone-induced changes in the metabolism of birch (Betula pendula) leaves. New Phytol. 137:
14             389-397.
15       Landolt, W.; Btihlmann, U.; Bleuler, P.; Bucher, J. B. (2000) Ozone exposure-response relationships for biomass and
16             root/shoot ratio of beech (Fagus sylvatica), ash (Fraxinus excelsior), Norway spruce (Picea abies) and scots
17             pine (Pinus sylvestris). Environ. Pollut. 109: 473-478.
18       Langebartels, C.; Kerner, K.; Leonardi, S.; Schraudner, M.; Trost, M.; Heller, W.;  Sandermann, H., Jr. (1991)
19             Biochemical plant responses to ozone: I. differential induction of polyamine and ethylene biosynthesis in
20             tobacco. Plant Physiol. 95: 882-889.
21       Langebartels, C.; Ernest, D.; Heller, W.; Lutz, C.; Payer, H.-D.; Sandermann, J. (1997) Ozone responses of trees:
22             results from controlled chamber exposures at the GSF phytotron. In: Sandermann, J. H.; Wellburn, A. R.; Heath,
23             R. L., eds. Forest decline and ozone: a comparison of controlled chamber and field experiments. Berlin,
24             Germany: Springer-Verlag; pp.  163-200. (Ecological studies: v.  127).
25       Larsen, R. I.; Heagle, A. S.; Heck, W. W. (1983) An air quality data analysis system for interrelating effects, standards,
26             and needed source reductions: part 7. an O3-SO2 leaf injury mathematical model. J. Air Pollut. Control Assoc.
27             33: 198-207.
28       Larsen, R. L; Heck, W.  W. (1984) An air quality data analysis system for interrelating effects, standards, and needed
29             source reductions: part 8. an effective mean O3 crop reduction mathematical model. J. Air Pollut. Control Assoc.
30             34: 1023-1034.
31       Larson, T. V.; Vong, R. J. (1990) A theoretical investigation of the pressure and temperature dependence of
32             atmospheric ozone  deposition to trees. Environ. Pollut. 67: 179-189.
33       Larson, J. L.; Zak, D. R.; Sinsabaugh, R. L. (2002) Extracellular enzyme activity beneath temperate trees growing
34             under elevated carbon dioxide and ozone. Soil Sci. Soc. Am. J. 66: 1848-1856.
35       Laurence, J. A.; Kohut, R. J.; Amundson, R. G.; Weinstein, D. A.; MacLean, D. C. (1996) Response of sugar maple to
36             multiple year exposure to ozone and simulated acidic precipitation. Environ. Pollut. 92: 119-126.
37       Laurence, J. A.; Amundson, R. G.; Kohut, R. J.; Weinstein, D. A. (1997) Growth and water use of red spruce (Picea
3 8             rubens Sarg.) exposed to ozone and simulated acidic precipitation for four growing seasons. For. Sci. 43:
39             355-361.
40       Laurence, J. A.; Retzlaff, W. A.; Kern, J. S.; Lee, E. H.; Hogsett,  W. E.; Weinstein, D.  A. (2001) Predicting the
41             regional impact of ozone and precipitation on the growth of loblolly pine and yellow poplar using linked
42             TREGRO and ZELG models. For. Ecol. Manage. 146: 247-263.
43       Law, M. Y.; Charles, S. A.; Halliwell, B. (1983) Glutathione and ascorbic acid in spinach (Spinacia oleracea)
44             chloroplasts: the  effect of hydrogen peroxide and of Paraquat. Biochem. J. 210: 899-903.
45       Lawlor, D. W. (1998) Plant responses to global change: temperature and drought stress. In: De Kok, L. J.; Stolen, I.,
46             eds. Responses of plant metabolism to air pollution and global change. Leiden, The Netherlands: Backhuys
47             Publishers; pp. 193-207.
48       Lawlor, D. W.; Keys, A. J. (1993) Understanding photosynthetic adaptation to changing climate. In: Fowden, L.;
49             Mansfield, T.; Stoddart, J., eds. Plant adaptation to environmental stress; London, United Kingdom: Chapman
50             and Hall; pp. 85-106.
51       Lawson, T.; Craigon, J.; Black, C. R.; Colls, J. J.; Tulloch, A.-M.; Landon, G. (2001a) Effects of elevated carbon
52             dioxide and ozone on the growth and yield of potatoes (Solanum tuberosum) grown in open-top chambers.
5 3             Environ.  Pollut. 111: 479-491.
         January 2005                                    9-371        DRAFT-DO NOT QUOTE OR CITE

-------
  1       Lawson, T.; Craigon, J.; Tulloch, A.-M.; Black, C. R.; Colls, J. I; Landon, G. (200Ib) Photosynthetic responses to
  2             elevated CO2 and O3 in field-grown potato (Solarium tuberosum). J. Plant Physiol. 158: 309-323.
  3       Lawson, T.; Craigon, J.; Black, C. R.; Colls, J. J.; Landon, G.; Weyers, J. D. B. (2002) Impact of elevated CO2 and O3
  4             on gas exchange parameters and epidermal characteristics in potato (Solarium tuberosum L.). J. Exp. Bot. 53:
  5             737-746.
  6       Le Thiec, D.; Manninen, S. (2003) Ozone and water deficit reduced growth of Aleppo pine seedlings. Plant Physiol.
  7             Biochem. 41: 55-63.
  8       Lee, E. H.; Bennett, J. H. (1982) Superoxide dismutase: a possible protective enzyme against ozone injury in snap
  9             beans (Phaseolus vulgaris L.). Plant Physiol. 69: 1444-1449.
10       Lee, E. H.; Hogsett, W. E. (1999) Role of concentrations and time of day in developing ozone exposure indices for a
11             secondary standard. J. Air Waste Manage. Assoc. 49: 669-681.
12       Lee, J. J.; Lewis, R. A.; Body, D. E. (1975) A field experimental system for the evaluation of the bioenvironmental
13             effects of sulfur dioxide. In: Proceedings of the Fort Union Coal Field Symposium; April; Billings, MT.
14             Billings, MT: Montana Academy of Science, v. 5, pp. 608-620.
15       Lee, E. H.; Tingey, D. T.; Hogsett, W. E. (1987) Selection of the best exposure-response model using various 7-hour
16             ozone exposure statistics. Research Triangle Park, NC: U.S. Environmental Protection Agency, Office of Air
17             Quality  Planning and Standards.
18       Lee, E. H.; Tingey, D. T.; Hogsett, W. E. (1988) Evaluation of ozone exposure indices in exposure-response modeling.
19             Environ. Pollut. 53: 43-62.
20       Lee, E. H.; Tingey, D. T.; Hogsett, W. E. (1989) Interrelation of experimental exposure and ambient air quality data
21             for comparison of ozone exposure indices and estimating agricultural losses. Corvallis, OR: U.S. Environmental
22             Protection Agency, Environmental Research Laboratory; EPA report no. EPA-600/3-89-047. Available from:
23             NTIS, Springfield, VA; PB89-195036.
24       Lee, E. H.; Hogsett, W. E.; Tingey, D. T. (1994a) Alternative attainment criteria for a secondary federal standard for
25             ozone. In: Nriagu, J. O.;  Simmons, M. S., eds. Environmental oxidants. New York, NY: John Wiley & Sons;
26             pp. 549-584. (Nriagu, J. O.,  ed. Advances in environmental  sciences and technology series: v. 28).
27       Lee, E. H.; Hogsett, W. E.; Tingey, D. T. (1994b) Attainment and effects issues regarding alternative secondary ozone
28             air quality standards.  J. Environ. Qual. 23: 1129-1140.
29       Lee, J. C.; Skelly, J. M.; Steiner, K. C.; Zhang, J. W.; Savage, J. E. (1999) Foliar response of black cherry (Prunus
30             serotina) clones to ambient ozone exposure in  central Pennsylvania. Environ. Pollut. 105: 325-331.
31       Lee, J. C.; Steiner, K. C.; Zhang, J. W.; Skelly, J. M.  (2002) Heritability of ozone sensitivity in open-pollinated
32             families of black cherry (Prunus serotina Ehrh.). For. Sci. 48: 111-117.
33       Lee, E. H.; Tingey, D. T.; Hogsett, W. E.; Laurence, J. A. (2003) History of tropospheric ozone for the San Bernardino
34             Mountains of southern California, 1963-1999.  Atmos. Environ. 37: 2705-2717.
3 5       Lefohn, A. S. (1984) A comparison of ambient ozone exposures for selected nonurban sites. Presented at: 77th annual
36             meeting of the Air Pollution Control Association; June; San Francisco, CA. Pittsburgh, PA: Air Pollution
37             Control  Association; paper no. 84-104.1.
38       Lefohn, A. S.; Benedict, H. M. (1982) Development of a mathematical index that describes ozone concentration,
39             frequency and duration. Atmos. Environ. 16: 2529-2532.
40       Lefohn, A. S.; Foley, J. K. (1992) NCLAN results and their application to the standard-setting process: protecting
41             vegetation from surface ozone exposures. J. Air Waste Manage. Assoc. 42:  1046-1052.
42       Lefohn, A. S.; Irving, P. M.  (1988) Characterizing ambient ozone exposure regimes in agricultural areas. Presented at:
43             81 st annual meeting of the Air Pollution Control Association; June; Dallas,  TX. Pittsburgh, PA: Air Pollution
44             Control  Association;  88-69.1.
45       Lefohn, A. S.; Jones, C. K.  (1986) The characterization of ozone and sulfur dioxide air quality data for assessing
46             possible vegetation effects. J. Air Pollut.  Control Assoc. 36: 1123-1129.
47       Lefohn, A. S.; Runeckles, V. C. (1987) Establishing standards to protect vegetation - ozone exposure/dose
48             considerations. Atmos. Environ. 21: 561-568.
49       Lefohn, A. S.;  Shadwick, D. S. (2000) Differences in trending estimates in the United States using several ozone
50             metrics. Presented at: 93rd annual meeting of the Air & Waste Management Association; June; Salt Lake City,
51             UT. Pittsburgh, PA: Air & Waste Management Association.
52       Lefohn, A. S.; Tingey, D. T. (1985) Comments on "Injury response of Phaseolus vulgaris to ozone flux density."
53             Atmos. Environ. 19: 206-207.
         January 2005                                    9-372        DRAFT-DO NOT QUOTE OR CITE

-------
  1       Lefohn, A. S.; Hogsett, W. E.; Tingey, D. T. (1986) A method for developing ozone exposures that mimic ambient
  2             conditions in agricultural areas. Atmos. Environ. 20: 361-366.
  3       Lefohn, A. S.; Davis, C. E.; Jones, C. K.; Tingey, D. T.; Hogsett, W. E. (1987) Co-occurrence patterns of gaseous air
  4             pollutant pairs at different minimum concentrations in the United States. Atmos. Environ. 21: 2435-2444.
  5       Lefohn, A. S.; Laurence, J. A.; Kohut, R. J. (1988a) A comparison of indices that describe the relationship between
  6             exposure to ozone and reduction in the yield of agricultural crops. Atmos. Environ. 22: 1229-1240.
  7       Lefohn, A. S.; Laurence, J. A.; Kohut, R. J. (1988b) A comparison of indices that describe the relationship between
  8             exposure to ozone and reduction in the yield of agricultural crops [response to comments by V. C. Runeckles].
  9             Atmos. Environ. 22: 1242-1243.
10       Lefohn, A. S.; Runeckles, V.  C.; Krupa, S. V.; Shadwick, D. S. (1989) Important considerations for establishing a
11             secondary ozone standard to protect vegetation. JAPCA 39: 1039-1045.
12       Lefohn, A. S.; Shadwick, D. S.; Mohnen, V. A. (1990) The characterization of ozone concentrations at a select set of
13             high-elevation sites in the eastern United States. Environ. Pollut. 67: 147-178.
14       Lefohn, A. S.; Shadwick, D. S.; Somerville, M. C.; Chappelka, A. H.; Lockaby, B. G.; Meldahl, R. S. (1992) The
15             characterization and comparision of ozone exposure indices used in assessing the response of loblolly pine to
16             ozone. Atmos. Environ. Part A 26: 287-298.
17       Lefohn, A. S.; Jackson, W.; Shadwick, D. S.; Knudsen, H. P. (1997) Effect of surface ozone exposures on vegetation
18             grown in the southern Appalachian Mountains: identification of possible areas of concern. Atmos. Environ. 31:
19             1695-1708.
20       Lefohn, A. S.; Shadwick, D. S.; Ziman, S. D. (1998) The difficult challenge of attaining EPA's new ozone standard.
21             Environ.  Sci. Technol. 32: 276A-282A.
22       Lefohn, A. S.; Oltmans, S. J.; Dann, T.; Singh, H. B. (2001) Present-day variability of background ozone in the lower
23             troposphere. J. Geophys. Res. [Atmos.] 106: 9945-9958.
24       Legge, A. H.; Griinhage, L.; Nosal, M.; Jager, H.-J.; Krupa, S. V. (1995) Ambient ozone and adverse crop response: an
25             evaluation of North American and European data as they relate to exposure indices and critical levels. Angew.
26             Bot. 69: 192-205.
27       Lendzian, K. J.; Kerstiens, G. (1991) Sorption and transport of gases and vapors in plant cuticles. Rev. Environ.
28             Contam. Toxicol. 121: 65-128.
29       Leon, J.; Lawton, M.  A.; Raskin, I. (1995) Hydrogen peroxide synthesis stimulates salicylic acid biosynthesis in
30             tobacco. Plant Physiol. 108: 1673-1683.
31       Lesser, V. M.; Rawlings, J. O.; Spruill, S. E.; Somerville, M. C. (1990) Ozone effects on agricultural crops: statistical
32             methodologies and estimated dose-response relationships. Crop Sci.  30: 148-155.
33       Levitt, J. (1972) Responses of plants to environmental stress. New York, NY: Academic Press.
34       Lichtenthaler, H. K. (1977) Regulation of prenylquinone synthesis in higher plants. In: Tevini, M.; Lichtenthaler, H.
35             K., eds. Lipids and lipid polymers in higher plants. New York, NY: Springer-Verlag; pp. 231-258.
36       Lindroth, R. L.; Reich, P. B.;  Tjoelker, M. G.; Volin, J. C.; Oleksyn, J. (1993) Light environment alters response to
37             ozone stress in seedlings of Acer saccharum March, and hybrid Populus L. III. Consequences for performance
38             of gypsy  moth. New Phytol. 124: 647-654.
39       Lindroth, R. L.; Kopper, B. J.; Parsons, W. F. J.; Bockheim, J. G.; Karnosky, D. F.; Hendrey, G. R.; Pregitzer, K. S.;
40             Isebrands, J. G.; Sober, J. (2001) Consequences of elevated carbon dioxide and ozone for foliar chemical
41             composition and dynamics in trembling aspen (Populus tremuloides) and paper birch (Betula papyri/era).
42             Environ.  Pollut. 115: 395-404.
43       Liu, L.;  Eriksson, K.-E. L.; Dean, J. F. D. (1995) Localization of hydrogen peroxide production in Pisum sativum L.
44             using epi-polarization microscopy to follow cerium perhydroxide disposition. Plant Physiol. 107: 501-506.
45       Lloyd, A.  C.; Lents, J. M.; Green, C.; Nemeth, P. (1989) Air quality management in Los Angeles: perspectives on past
46             and future emission control strategies. JAPCA 39: 696-703.
47       Loats, K. V.; Rebbeck, J. (1999) Interactive effects of ozone and elevated carbon dioxide on the growth and
48             physiology of black cherry, green ash, and yellow poplar seedlings. Environ. Pollut. 106: 237-248.
49       Logan, J. A. (1989) Ozone in rural areas of the United States. J. Geophys. Res. [Atmos.] 94: 8511-8532.
50       Long, S. P. (1991) Modification of the response of photosynthetic productivity to rising temperature by atmospheric
51             CO2 concentrations: has its importance been underestimated? Plant Cell Environ. 14: 729-739.
52       Longstreth, J. D.; De  Gruijl; Kripke, M. L.; Takizawa, Y.; van der Leun, J.  C. (1995) Effects of increased solar
53             ultraviolet radiation on human health. Ambio 24: 153-165.
         January 2005                                    9-373        DRAFT-DO NOT QUOTE OR CITE

-------
  1       Lorenzini, G.; Nali, C. (1995) Analysis of vertical ozone and nitrogen oxides profiles in a Primus cerasifera canopy.
  2             Int. J. Biometeorol. 39:1-4.
  3       Lorenzini, G.; Medeghini Bonatti, P.; Nali, C.; Baroni Fornasiero, R. (1994) The protective effect of rust infection
  4             against ozone, sulphur dioxide and paraquat toxicity symptoms in broad bean. Physiol. Mol. Plant Pathol. 45:
  5             263-279.
  6       Lucas, P. W.;  Cottam, D. A.; Mansfield, T. A. (1987) A large-scale fumigation system for investigating interactions
  7             between air pollution and cold stress on plants. Environ. Pollut. 43: 15-28.
  8       Lucas, P. W.;  Cottam, D. A.; Sheppard, L. J.; Francis, B. J. (1988) Growth responses and delayed winter hardening in
  9             Sitka spruce following summer exposure to ozone. New Phytol. 108: 495-504.
10       Luo, Y. (2001) Transient ecosystem response to free-air CO2 enrichment (FACE): experimental evidence and methods
11             of analysis. New Phytol. 152: 3-8.
12       Luo, Y.; Reynolds, J. F. (1999) Validity of extrapolating field CO2 experiments to predict carbon sequestration in
13             natural ecosystems. Ecology 80: 1568-1583.
14       Luo, Y.; Canadell, J.; Mooney, H. A.  (1999) Interactive effects of carbon dioxide and environmental stress on plants
15             and ecosystems: a synthesis. In: Luo, Y.; Mooney, H. A., eds. Carbon dioxide and environmental stress. San
16             Diego,  CA: Academic Press; pp. 393-408.
17       Lutz, C.;  Anegg, S.; Gerant, D.; Alaoi-Sosse, B.; Gerard, J.; Dizengremel, P. (2000) Beech trees exposed to high CO2
18             and to simulated summer ozone levels: effects on photosynthesis, chloroplast components and leaf enzyme
19             activity. Physiol. Plant. 109: 252-259.
20       Luwe, M. W. F.; Takahama, U.; Heber, U. (1993) Role of ascorbate in detoxifying ozone in the apoplast of spinach
21             (Spinacia oleracea L.) leaves. Plant Physiol. 101: 969-976.
22       Luxmoore, R. J. (1988) Assessing the mechanisms of crop loss from air pollutants with process models. In: Heck, W.
23             W.; Taylor, O.  C.; Tingey, D. T., eds. Assessment of crop loss from air pollutants: proceedings of an
24             international conference; October 1987; Raleigh, NC. Essex, United Kingdom: Elsevier Science Publishers,
25             Ltd.; pp. 417-444.
26       Lyons, T. M.; Barnes, J. D. (1998) Influence of plant age on ozone resistance inPlantago major. New Phytol. 138:
27             83-89.
28       Lyons, T. M.; Barnes, J. D.; Davison, A. W. (1997) Relationships between ozone resistance and climate in European
29             populations ofPlantago major. New Phytol. 136: 503-510.
30       Lyons, T.; Ollerenshaw, J. H.; Barnes, J. D. (1999) Impacts of ozone onPlantago major: apoplastic and symplastic
31             antioxidant status. New Phytol. 141: 253-263.
32       MacLean, D. C. (1990) Joint action of ozone and hydrogen fluoride on foliar senescence in maize. Environ. Pollut. 63:
33             283-292.
34       Madamanchi,  N. R.; Anderson, J. V.; Alscher, R. G.; Cramer, C. L.; Hess, J. L. (1992) Purification of multiple forms
35             of glutathione reductase from pea (Pisum sativum L.) seedlings and enzyme levels in ozone-fumigated pea
36             leaves.  Plant Physiol. 100: 138-145.
37       Magan, N.; Kirkwood, I. A.; McLeod, A. R.; Smith, M. K. (1995) Effect of open-air fumigation with sulphur dioxide
38             and ozone on phyllosphere and endophytic fungi of conifer needles. Plant Cell Environ. 18: 291-302.
39       Maggs, R.; Ashmore, M. R. (1998) Growth and yield responses of Pakistan rice (Oryza sativa L.) cultivars to O3 and
40             NO2. Environ. Pollut. 103: 159-170.
41       Maier-Maercker, U. (1997) Experiments on the water balance of individual attached twigs ofPicea abies (L.) Karst. in
42             pure and ozone-enriched air. Trees 11: 229-239.
43       Maier-Maercker, U. (1998) Predisposition of trees to drought stress by ozone. Tree Physiol. 19: 71-78.
44       Maier-Maercker, U.; Koch, W. (1992) The effect of air pollution on the mechanism of stomatal control. Trees (Berlin)
45             7:  12-25.
46       Manderscheid, R.; Jager, H.-J.; Kress, L. W. (1992) Effects of ozone on foliar nitrogen metabolism of Pinus taedaL.
47             and implications for carbohydrate metabolism. New Phytol. 121: 623-633.
48       Mandl, R. H.; Weinstein, L. H.; McCune, D. C.; Keveny, M. (1973) A cylindrical, open-top chamber for the exposure
49             of plants to air pollutants in the field. J. Environ. Qual. 2: 371-376.
50       Mandl, R. H.; Laurence, J. A.; Kohut, R. J.  (1989) Development and testing of open-top  chambers for exposing large,
51             perennial plants to air pollutants. J. Environ. Qual. 18: 534-540.
52       Manninen, A.-M.; Holopainen, T.; Lyytikainen-Saarenmaa, P.; Holopainen, J. K. (2000) The  role of low-level ozone
53             exposure and mycorrhizas in chemical quality and insect herbivore performance on Scots pine seedlings. Global
54             Change Biol. 6: 111-121.


         January 2005                                    9-374        DRAFT-DO NOT QUOTE OR CITE

-------
  1       Manninen, S.; Siivonen, N.; Timonen, U.; Huttunen, S. (2003) Differences in ozone response between two Finnish
  2             wild strawberry populations. Environ. Exp. Bot. 49: 29-39.
  3       Manning, W. J. (2000) Use of protective chemicals to  assess the effects of ambient ozone on plants. In: Agrawal, S. B.;
  4             Agrawal, M, eds. Environmental pollution and plant responses. Boca Raton, FL: Lewis Publishers; pp.
  5             247-258.
  6       Manning, W. J. (2003) Detecting plant effects is necessary to give biological significance to ambient ozone monitoring
  7             data and predictive ozone standards. Environ. Pollut. 126: 375-379.
  8       Manning, W. J.; Keane, K. D. (1988) Effects of air pollutants on interactions between plants, insects, and pathogens.
  9             In: Heck, W. W.; Taylor, O. C.; Tingey, D. T., eds. Assessment of crop loss from air pollutants: proceedings of
10             an international conference; October 1987; Raleigh, NC. Essex, United Kingdom: Elsevier Science Publishers,
11             Ltd.; pp. 365-386.
12       Manning, W. J.; Krupa, S. V. (1992) Experimental methodology for studying the effects of ozone on crops and trees.
13             In: Lefohn, A.  S., ed. Surface level ozone exposures and their effects on vegetation. Chelsea, MI: Lewis
14             Publishers, Inc.; pp. 93-156.
15       Manning, W. J.; Feder, W. A.; Vardaro, P. M. (1974) Suppression of oxidant injury by benomyl: effects on yields of
16             bean cultivars in the field. J. Environ. Qual. 3: 1-3.
17       Manning, W. J.; Godzik, B.; Musselman, R. (2002) Potential bioindicator plant species for ambient ozone in forested
18             mountain areas of central Europe. Environ. Pollut. 119: 283-290.
19       Manning, W. J.; Flagler, R. B.; Frenkel, M. A. (2003)  Assessing plant response to ambient ozone: growth of
20             ozone-sensitive loblolly pine seedlings treated with ethylenediurea or sodium erythorbate. Environ. Pollut. 126:
21             73-81.
22       Martin, M. J.; Host, G. E.; Lenz, K. E.;  Isebrands, J. G. (2001) Simulating the growth response of aspen to elevated
23             ozone: a mechanistic approach to scaling a leaf-level model of ozone effects on photosynthesis to a complex
24             canopy architecture. Environ. Pollut. 115: 425-436.
25       Massman, W. J. (1993) Partitioning ozone fluxes to sparse grass and soils and the inferred resistances to dry
26             deposition. Atmos. Environ. Part A 27A: 167-174.
27       Massman, W. J. (2004) Toward an ozone standard to protect vegetation based on effective dose: a review of deposition
28             resistances and a possible metric. Atmos. Environ. 38: 2323-2337.
29       Massman, W. J.; Grantz, D. A. (1995) Estimating canopy conductance to ozone uptake from observations of
30             evapotranspiration at the canopy scale and at the leaf scale. Global Change Biol. 1: 183-198.
31       Massman, W. J.; Ham, J. M. (1994) An evaluation of a surface-energy balance method for partitioning ET data into
32             plant and soil components for a surface with partial canopy cover. Agric. For. Meteorol. 67: 253-267.
33       Massman, W. J.; Pederson, J.; Delany, A.; Grantz, D.; Denhartog, G.; Neumann, H. H.; Oncley, S. P.; Pearson, R.;
34             Shaw, R. H. (1994) An evaluation of the regional acid deposition model surface module for ozone uptake at 3
35             sites in the San-Joaquin valley of California. J. Geophys. Res. [Atmos.] 99: 8281-8294.
36       Massman, W. J.; Musselman, R. C.; Lefohn, A. S. (2000) A conceptual ozone dose-response model to develop a
37             standard to protect vegetation. Atmos. Environ. 34: 745-759.
38       Matters, G. L.; Scandalios, J. G. (1987) Synthesis of isozymes of superoxide dismutase in maize leaves in response to
39             O3, SO2, and elevated O2. J. Exp. Bot. 38:  842-852.
40       Matyssek, R.; Gunthardt-Goerg, M. S.;  Maurer, S.; Keller, T. (1995) Nighttime exposure to ozone reduces whole-plant
41             production inBetulapendula. Tree Physiol. 15: 159-165.
42       Matyssek, R.; Gtinthardt-Goerg, M. S.;  Maurer, S.; Christ, R. (2002) Tissue structure and respiration of stems of
43             Betulapendula under contrasting ozone exposure and nutrition. Trees 16: 375-385.
44       Maurer, S.; Matyssek, R. (1997) Nutrition and the ozone sensitivity of birch (Betulapendula) II. Carbon balance,
45             water-use  efficiency and nutritional status of the whole plant. Trees 12: 11-20.
46       Mazarura, U. (1997) Effects of fluctuating levels of ozone and nitrogen dioxide alone and in  sequence on plant growth
47             and photosynthesis [dissertation]. Vancouver, British Columbia, Canada: University of British Columbia.
48       McBride, J. R.; Laven, R. D. (1976) Scars as an indicator of fire frequency in the San Bernardino Mountains,
49             California. J. For. 74: 439-442.
50       McBride, J. R.; Laven, R. D. (1999) Impact of oxidant air pollutants on forest succession in the mixed conifer forests
51             of the San Bernardino Mountains. In: Miller, P. R.; McBride, J. E., eds. Oxidant air pollution impacts in the
52             montane forests of southern California: a case study of the San Bernardino Mountains. New York, NY:
53             Springer; pp. 338-352. (Ecological Studies, v. 134).
         January 2005                                    9-375        DRAFT-DO NOT QUOTE OR CITE

-------
  1       McBride, J. R.; Semion, V. P.; Miller, P. R. (1975) Impact of air pollution on the growth of Ponderosa pine. Calif.
  2             Agric. 29: 8-9.
  3       McCarl, B. A.; Alig, R.; Adams, D. M. (1998) Impacts of incorporating land exchanges between forestry and
  4             agriculture in sector models. J. Agric. Appl. Econ. 30: 389-401.
  5       McCool, P. M.; Musselman, R. C.; Younglove, T.; Teso, R. R. (1988) Response of kidney bean to sequential ozone
  6             exposures. Environ. Exp. Bot. 28: 307-313.
  7       McDonald, E. P.; Kruger, E. L.; Riemenschneider, D. E.; Isebrands, J. G. (2002) Competitive status influences
  8             tree-growth responses to elevated CO2 and O3 in aggrading aspen stands. Funct. Ecol. 16: 792-801.
  9       McGartland, A. M. (1987) The implications of ambient ozone standards for U.S. agriculture: a comment and some
10             further evidence. J. Environ. Manage. 24: 139-146.
11       McKee, I. F.; Farage, P. K.; Long, S. P. (1995) The interactive effects of elevated CO2 and O3 concentration on
12             photosynthesis in spring wheat. Photosyn. Res. 45: 111-119.
13       McKee, I. F.; Bullimore, J. F.; Long, S. P. (1997a) Will elevated CO2 concentrations protect the yield of wheat from O3
14             damage? Plant Cell Environ. 20: 77-84.
15       McKee, I. F.; Eiblmeier, M.; Polle, A. (1997b) Enhanced ozone-tolerance in wheat grown at an elevated CO2
16             concentration: ozone exclusion and detoxification. New Phytol.  137: 275-284.
17       McKee, I. F.; Mulholland, B.  J.; Craigon, J.; Black, C. R.; Long, S. P. (2000) Elevated concentrations of atmospheric
18             CO2 protect against and compensate for O3 damage to photosynthetic tissues of field-grown wheat. New Phytol.
19             146:427-435.
20       McLaughlin, S. B.; Downing, D. J. (1995) Interactive effects of ambient ozone and climate measured on growth of
21             mature forest trees. Nature (London) 374: 252-254.
22       McLaughlin, S. B.; Downing  , D. J. (1996) Interactive effects of ambient ozone and climate measured on growth of
23             mature loblolly pine trees. Can. J. For. Res. 670-681.
24       McLaughlin, S.; Percy, K. (1999) Forest health in North America: some perspectives on actual and potential roles of
25             climate and air pollution. Water Air Soil Pollut. 116:151-197.
26       McLaughlin, S. B.; McConathy, R. K.; Duvick, D.; Mann, L. K. (1982) Effects of chronic air pollution stress on
27             photosynthesis, carbon allocation, and growth of white pine trees. For. Sci. 28: 60-70.
28       McLaughlin, S. B.; Layton, P. A.;  Adams, M. B.; Edwards, N. T.; Hanson, P. J.; O'Neill, E. G.; Roy, W. K. (1994)
29             Growth-responses of 53 open-pollinated loblolly-pine families to ozone  and acid-rain. J. Environ. Qual. 23:
30             247-257.
31       McLaughlin, S. B.; Shortle, W. C.; Smith, K. T. (2002) Dendroecological applications in air pollution and
32             environmental chemistry: research needs. Dendrochronologia 20: 133-157.
33       McLaughlin, S. B.; Wullschleger,  S. D.; Nosal, M. (2003) Diurnal and seasonal changes in stem increment and water
34             use by yellow poplar trees in response to environmental stress. Tree Physiol. 23: 1125-1136.
35       McLeod, A. R.; Fackrell, J. E.; Alexander, K. (1985) Open-air fumigation of field crops: criteria and design for a new
36             experimental system. Atmos. Environ. 19: 1639-1649.
37       McLeod, A. R.; Shaw, P. J. A.; Holland, M. R. (1992) The Liphook Forest Fumigation Project: studies of sulphur
38             dioxide and ozone effects on coniferous trees. For. Ecol. Manage. 51: 121-127.
39       Mearns, L. O.; Rosenzweig, C.; Goldberg, R. (1997) Mean and variance change in climate scenarios:  methods,
40             agricultural applications and measures of uncertainty. Clim. Change 35:  367-396.
41       Mehdy, M. C. (1994) Active oxygen species in plant defense against pathogens. Plant Physiol. 105: 467-472.
42       Mehlhorn, H.; Wellburn, A. R. (1987) Stress ethylene formation determines plant sensitivity to ozone. Nature
43             (London) 327: 417-418.
44       Mehlhorn, H.; Wenzel, A. A.  (1996) Manganese deficiency enhances ozone toxicity in bush beans  (Phaseolus vulgaris
45             L. cv. Saxa). J. Plant Physiol. 148: 155-159.
46       Mehlhorn, H.; Tabner, B. J.; Wellburn, A. R. (1990) Electron spin resonance evidence for the formation of free
47             radicals in plants exposed to ozone. Physiol. Plant. 79: 377-383.
48       Mehlhorn, H.; O'Shea, J. M.; Wellburn, A. R. (1991) Atmospheric ozone interacts with stress ethylene formation by
49             plants to cause visible plant injury. J. Exp. Bot. 42: 17-24.
50       Meinhard, M.; Rodriguez, P. L.; Grill, E. (2002) The sensitivity of ABI2 to hydrogen peroxide links the abscisic
51             acid-response regulator to redox signalling. Planta 214: 775-782.
52       Mendelsohn, R.; Nordhaus, W.; Shaw, D. (1994) The impact of global warming on agriculture: a Ricardian analysis.
53             Am. Econ. Rev. 84: 753-771.
         January 2005                                   9-376        DRAFT-DO NOT QUOTE OR CITE

-------
  1       Miller, P. L. (1973) Oxidant-induced community change in a mixed conifer forest. In: Naegele, J. A., ed. Air pollution
  2             damage to vegetation. Washington, DC: American Chemical Society; pp. 101-117. (Advances in chemistry
  3             series: v. 122).
  4       Miller, P. R.; Elderman, M. J. (1977) Photochemical oxidant air pollutant effects on a mixed conifer forest ecosystem:
  5             a progress report, 1976. Corvallis, OR: U.S. Environmental Protection Agency, Corvallis Environmental
  6             Research Laboratory; EPA report no. EPA-600/3-77-104. Available from: NTIS, Springfield, VA; PB-274531.
  7       Miller, P. R.; McBride, J. R. (1999) Assessment of ecological risks and implications for policy and management in the
  8             San Bernardino Mountains. In: Miller, P. R.; McBride, J. R., eds. Oxidant air pollution impacts in the Montane
  9             forests of southern California: a case study of the San Bernardino Mountains. New York, NY: Springer; pp.
10             397-404. (Ecological Studies, v.  134).
11       Miller, P. R.; Millecan, A. A. (1971) Extent of oxidant air pollution damage to some pines and other conifers in
12             California. Plant Dis. Rep.  55: 555-559.
13       Miller, P. R.; Rechel, J. (1999) Temporal changes in crown condition indices, needle litterfall, and collateral needle
14             injuries of Ponderosa and Jeffrey pines. In: Miller, P. R.; McBride, J. R., eds. Oxidant air pollution impacts in
15             the Montane forests of southern California: a case study of the San Bernardino Mountains. New York, NY:
16             Springer; pp.  164-178.
17       Miller, P. R.; Parmeter, J. R., Jr.; Taylor, O. C.; Cardiff, E. A. (1963) Ozone injury to the foliage ofPinuspondemsa.
18             Phytopathology 53: 1072-1076.
19       Miller, P. R.; Parmeter, J. R., Jr.; Flick, B. H.; Martinez, C. W. (1969) Ozone dosage response of ponderosa pine
20             seedlings. J. Air Pollut. Control Assoc.  19: 435-438.
21       Miller, P. R.; McCutchan, M. H.; Milligan, H. P. (1972) Oxidant air pollution in the Central Valley,  Sierra Nevada
22             foothills, and Mineral King Valley of California. Atmos. Environ. 6: 623-633.
23       Miller, P. R.; McBride, J. R.; Schilling, S. L.; Gomez, A. P. (1989) Trend of ozone damage to conifer forests between
24             1974 and 1988 in the San Bernardino Mountains of southern California. In: Olson, R. K.; Lefohn, A. S., eds.
25             Effects of air pollution on western forests [an A&WMA symposium; June; Anaheim, CA]. Pittsburgh, PA: Air
26             and Waste Management Association; pp. 309-323. (APCA transactions series: no. 16).
27       Miller, J. E.; Booker, F. L.; Fiscus, E. L.; Heagle, A. S.; Pursley, W. A.; Vozzo, S. F.; Heck, W. W. (1994)
28             Ultraviolet-B radiation and ozone effects on growth, yield, and photosynthesis of soybean. J. Environ. Qual. 23:
29             83-91.
30       Miller, J. E.; Heagle, A. S.; Pursley, W. A. (1998) Influence of ozone stress on soybean response to carbon dioxide
31             enrichment: II. Biomass and development. Crop Sci. 38: 122-128.
32       Miller, P. R.; Bytnerowicz, A.; Fenn, M.; Poth, M.; Temple, P.; Schilling, S.; Jones, D.; Johnson, D.; Chow, J.;
33             Watson, J. (1998) Multidisciplinary study of ozone, acidic deposition and climatic effects on a mixed conifer
34             forest in California, USA. Chemosphere 36: 1001-1006.
35       Mills, G. (2002) Modification of plant response by environmental conditions. In: Bell, J. N. B.; Treshow, M., eds. Air
36             Pollution and Plant Life. 2nd ed. Chichester, United Kingdom: John Wiley & Sons Ltd.; pp. 343-358.
37       Mills, G.; Ball, G.; Hayes, F.; Fuhrer, J.; Skarby, L.; Gimeno, B.; De Temmerman, L.; Heagle, A.; Members of the ICP
3 8             Vegetation programme. (2000) Development of a multi-factor model for predicting the effects of ambient ozone
39             on the biomass of white clover. Environ. Pollut. 109: 533-542.
40       Minnich, R. A. (1999) Vegetation, fire regimes, and forest dynamics. In: Oxidant air pollution impacts in the montane
41             forests southern California: a case study of the San Bernardino mountains. New York, NY: Springer; pp. 44-80.
42             (Ecological studies: v. 134).
43       Minnich, R. A.; Padgett, P. E. (2003) Geology, climate, and vegetation of the Sierra Nevada and the mixed-conifer
44             zone: an introduction to the ecosystem.  In: Bytnerowicz, A.; Arbaugh, M. J.; Alonso, R., eds. Ozone Air
45             Pollution in the Sierra Nevada: Distribution and Effects on Forests. New York, NY: Elsevier; pp. 1-31.
46             (Developments in Environmental Science, v. 2).
47       Minnich, R. A.; Barbour, M. G.; Burk, J. H.; Fernau, R. F. (1995) Sixty years of change in California conifer forests of
48             the San Bernardino mountains. Conserv. Biol. 9: 902-914.
49       Moldau, H. (1998) Hierarchy of ozone scavenging reactions in the plant cell wall. Physiol. Plant. 104: 617-622.
50       Moldau, H.; Bichele, I. (2002) Plasmalemma protection by the apoplast as assessed from above zero ozone
51             concentration in leaf intercellular air spaces. Planta 214:  484-487.
52       Momen, B.; Helms,  J. A. (1996) Effects of simulated acid rain and ozone on foliar chemistry of field-grown Pinus
53             ponderosa seedlings and mature  trees. Environ. Pollut. 91: 105-111.
         January 2005                                    9-377        DRAFT-DO NOT QUOTE OR CITE

-------
  1       Momen, B.; Helms, J. A.; Criddle, R. S. (1996) Foliar metabolic heat rate of seedlings and mature trees of Firms
  2             ponderosa exposed to simulated acid rain and ozone. Plant Cell Environ. 19: 747-753.
  3       Momen, B.; Anderson, P. D.; Helms, J. A.; Houpis, J.  L. J. (1997) Acid rain and ozone effects on gas exchange of
  4             Pinusponderosa'. a comparison between trees and seedlings. Int. J. Plant Sci. 158: 617-621.
  5       Momen, B.; Anderson, P. D.; Helms, J. A. (1999) Temperature dependency of acid-rain effect on photosynthesis of
  6             Pinus ponderosa. For. Ecol. Manage. 113: 223-230.
  7       Momen, B.; Anderson, P. D.; Houpis, J. L. J.; Helms, J. A. (2002) Growth of ponderosa pine seedlings as affected by
  8             air pollution. Atmos. Environ. 36: 1875-1882.
  9       Monteith, J. L.; Elston, J. (1993) Climatic constraints on crop production. In: Fowden, L.; Mansfield, T.; Stoddart, J.,
10             eds. Plant adaptation to environmental stress. London, United Kingdom: Chapman and Hall; pp. 3-18.
11       Morgan, P. B.; Ainsworth, E. A.; Long, S. P. (2003) How does elevated ozone impact soybean? A meta-analysis of
12             photosynthesis, growth and yield. Plant Cell Environ. 26:  1317-1328.
13       Morison, J. I. L.; Lawlor, D. W. (1999) Interactions between increasing CO2 concentration and temperature on plant
14             growth. Plant Cell Environ. 22: 659-682.
15       Mortensen, L. M. (1998) Growth responses of seedlings of six Betula pubescens Ehrh. Provenances to six ozone
16             exposure regimes. Scand. J. For. Res. 13: 189-196.
17       Mortensen, L. M. (1999) Foliar injuries caused by ozone in Betula pubescens Ehrh. and Phleum pratense L. as
18             influenced by climatic conditions before and during O3 exposure. Acta Agric. Scand. B 49: 44-49.
19       Mortensen, L.; Bastrup-Birk, A. (1996) Critical levels for O3 for biomass production in white clover. In: Karenlampi,
20             L.; Skarby, L., eds. Critical levels for ozone in Europe: testing and finalizing the concepts: UN-ECE workshop
21             report. UN-ECE Convention on Long-Range Transboundary Air Pollution; April; Kuopio, Finland. Kuopio,
22             Finland: University of Kuopio, Department of Ecology and Environmental Science; pp. 285-289.
23       Mudd, J. B. (1996) Biochemical basis for the toxicity of ozone. In: Yunus, M.; Iqbal, M., eds. Plant response to air
24             pollution. New York, NY: John Wiley & Sons; pp. 267-283.
25       Mudd, J. B.; Kozlowski, T.  T., eds. (1975) Responses  of plants to air pollution. New York, NY: Academic Press.
26             (Physiological ecology, a series of monographs, texts, and treatises).
27       Mudd, J. B.; Leavitt, R.; Ongun, A.; McManus, T. T. (1969) Reaction of ozone with amino acids and proteins. Atmos.
28             Environ. 3: 669-681.
29       Mudd, J. B.; McManus, T. T.; Ongun, A.  (1971) Inhibition of lipid metabolism in chloroplasts by ozone. In: Englund,
30             H. M.; Beery, W. T., eds. Proceedings of the second clean air congress. New York, NY: Academic Press; pp.
31             256-260.
32       Mudd, J. B.; Dawson, P. J.; Adams, J. R.; Wingo, J.; Santrock, J. (1996) Reaction of ozone with enzymes of
33             erythrocyte membranes. Arch. Biochem. Biophys. 335: 145-151.
34       Mudd, J. B.; Dawson, P. J.; Santrock, J. (1997a) Ozone does not react with human erythrocyte membrane lipids. Arch.
35             Biochem. Biophys. 341:251-258.
36       Mudd, J. B.; Dawson, P. J.; Tseng, S.; Liu, F.-P. (1997b) Reaction of ozone with protein tryptophans: band III, serum
37             albumin, and cytochrome C. Arch.  Biochem. Biophys. 338: 143-149.
3 8       Mueller, S. F. (1994) Characterization of ambient ozone levels in the Great Smoky Mountains National Park. J. Appl.
39             Meteorol. 33:465-472.
40       Mukammal, E. I.  (1965) Ozone as a cause of tobacco injury. Agric. Meteorol. 2: 145-165.
41       Mulchi, C. L.; Slaughter, L.; Saleem, M.; Lee, E. H.; Pausch, R.; Rowland, R. (1992) Growth and physiological
42             characteristics of soybean in open-top chambers in response to ozone and increased atmospheric CO2. Agric.
43             Ecosyst. Environ. 38: 107-118.
44       Mulchi, C.;Rudorff, B.; Lee, E.; Rowland, R.; Pausch, R. (1995) Morphological responses among crop species to
45             full-season exposures to enhanced concentrations of atmospheric CO2 and O3. Water Air Soil Pollut. 85:
46             1379-1386.
47       Mulholland, B.; Craigon, J.; Black, C.;  Atherton, J.; Landon, G. (1997a) Effects of elevated carbon dioxide and ozone
48             on the growth and yield of spring wheat (Triticum aestivum L.). J. Exper. Bot.  48: 113-122.
49       Mulholland, B. J.; Craigon, J.; Black, C. R.; Colls, J. J.; Atherton, J.; Landon, G. (1997b) Impact of elevated
50             atmospheric CO2 and O3 on gas exchange and chlorophyll content in spring wheat (Triticum aestivum L.). J.
51             Exp. Bot. 48: 1853-1863.
52       Mulholland, B. J.; Craigon, J.; Black, C. R.; Colls, J. J.; Atherton, J.; Landon, G. (1998a) Effects of elevated CO2 and
53             O3 on the rate and duration of grain growth and harvest index in spring wheat (Triticum aestivum L.). Global
54             Change Biol. 4: 627-635.


         January 2005                                    9-378       DRAFT-DO NOT QUOTE OR CITE

-------
  1       Mulholland, B. I; Craigon, I; Black, C. R.; Colls, J. I; Atherton, I; Landon, G. (1998b) Growth, light interception
  2             and yield responses of spring wheat (Triticum aestivum L.) grown under elevated CO2 and O3 in open-top
  3             chambers. Global Change Biol. 4: 121-130.
  4       Muntifering, R. B.; Crosby, D. D.; Powell, M. C.; Chappelka, A. H. (2000) Yield and quality characteristics of
  5             bahiagrass (Paspalum notatum) exposed to ground-level ozone. Anim. Feed Sci. Technol. 84: 243-256.
  6       Murphy, J. J.; Deluki, M. A.; McCubbin, D. R.; Kim, H. J. (1999) The cost of crop damage caused by ozone air
  7             pollution from motor vehicles. J. Environ. Manage. 55: 273-289.
  8       Musselman, R. C.; Hale, B. A. (1997) Methods for controlled and field ozone exposures of forest tree species in North
  9             America. In: Sandermann, H.; Wellburn, A. R.; Health, R. L., eds. Forest decline and ozone. Berlin, Germany:
10             Springer-Verlag; pp. 277-315.  (Ecological Studies, v. 127).
11       Musselman, R. C.; Massman, W. J. (1999) Ozone flux to vegetation and its relationship to plant response and ambient
12             air quality standards. Atmos. Environ. 33: 65-73.
13       Musselman, R. C.; Minnick, T. J. (2000) Nocturnal stomatal conductance and ambient air quality standards for ozone.
14             Atmos. Environ. 34: 719-733.
15       Musselman, R. C.; Oshima, R. J.; Gallavan, R. E. (1983) Significance of pollutant concentration distribution in the
16             response of'red kidney'beans to ozone. J. Am. Soc. Hortic. Sci. 108: 347-351.
17       Musselman, R. C.; Huerta, A. J.; McCool, P. M.; Oshima, R.  J. (1986) Response of beans to simulated ambient and
18             uniform ozone distributions with equal peak concentration. J. Am. Soc. Hortic. Sci. Ill: 470-473.
19       Musselman, R. C.; McCool, P. M.; Younglove, T. (1988) Selecting ozone exposure statistics for determining crop
20             yield loss from air pollutants. Environ. Pollut. 53: 63-78.
21       Musselman, R. C.; Younglove, T.; McCool, P. M. (1994) Response ofPhaseolus vulgaris L. to differing ozone
22             regimes having indentical total exposure. Atmos. Environ. 28: 2727-2731.
23       Neill, S. J.; Desikan, R.; Clarke, A.; Hurst, R. D.; Hancock, J. T. (2002) Hydrogen peroxide and nitric oxide as
24             signalling molecules in plants. J. Exper. Bot. 53:  1237-1247.
25       Neill, S.; Desikan, R.; Hancock, J. (2002) Hydrogen peroxide signalling. Curr. Opin. Plant Biol. 5: 388-395.
26       Neufeld, H. S.; Lee, E. H.; Renfro, J. R.; Hacker, W. D. (2000) Seedling insensitivity to ozone for three conifer species
27             native to Great Smoky Mountains National Park. Environ. Pollut. 108: 141-151.
28       Neufeld, H. S.; Renfro, J. R.; Hacker, W. D.; Silsbee, D. (1992) Ozone in Great Smoky Mountains National Park:
29             dynamics and effects on plants. In:  Berglund, R. L., ed. Tropospheric ozone and the environment II: effects,
30             modeling and control [papers from an international specialty conference];  November; Atlanta, GA. Pittsburgh,
31             PA: Air & Waste Management Association; pp. 594-617. (A&WMA transactions series no.  20).
32       Neufeld, H. S.; Lee, E. H.; Renfro, J. R.; Hacker, W. D.; Yu,  B.-H. (1995) Sensitivity of seedlings of black cherry
33             (Primus semtina Ehrh.) to ozone in Great Smoky Mountains National Park. I. Exposure-response curves for
34             biomass. New Phytol. 130: 447-459.
35       Niewiadomska, E.; Gaucher-Veilleux, C.; Chevrier, N.; Mauffette, Y.; Dizengremel, P. (1999) Elevated CO2 does not
36             provide protection against ozone considering the activity of several antioxidant enzymes in the leaves of sugar
37             maple. J. Plant Physiol. 155: 70-77.
38       Nijs, I.; Teughels, H.; Blum, H.; Hendrey, G.; Impens, I. (1996) Simulation of climate change with infrared heaters
39             reduces the productivity of Loliumperenne L. in summer. Environ. Exp. Bot. 36: 271-280.
40       Nikolov, N.; Zeller, K. (2003) Modeling coupled interactions of carbon, water, and ozone exchange between terrestrial
41             ecosystems and the atmosphere. I. Model description. Environ. Pollut. 124: 231-246.
42       Noormets, A.; McDonald, E. P.; Dickson, R. E.; Kruger, E. L.; Sober, A.; Isebrands, J.  G.; Karnosky, D. F. (2001) The
43             effect of elevated carbon dioxide and ozone on leaf- and branch-level photosynthesis and potential plant-level
44             carbon gain in aspen. Trees 15: 262-270.
45       Noormets, A.; Sober, A.; Pell, E. J.; Dickson, R. E.; Posila, G. K.; Sober, J.; Isebrands,  J. G.; Karnosky, D. F. (2001)
46             Stomatal and non-stomatal limitation to photosynthesis in two trembling aspen (Populus tremuloides Michx.)
47             clones exposed to elevated CO2 and/or O3. Plant Cell Environ. 24: 327-336.
48       Norby, R. J.; Wullschleger, S. D.; Gunderson, C. A.; Johnson, D. W.; Ceulemans, R. (1999) Tree responses to rising
49             CO2 in field experiments: implications for the future forest. Plant Cell Environ. 22: 683-714.
50       Norton, B. G. (1987) Why preserve natural variety? Princeton, NJ: Princeton University Press.
51       Nouchi, I.; Aoki, K. (1979) Morning glory as a photochemical oxidant indicator.  Environ. Pollut. 18: 289-303.
52       Nouchi, I.; Toyama, S. (1988) Effects of ozone and peroxyacetyl nitrate on polar lipids  and fatty acids in leaves of
53             morning glory and kidney bean. Plant Physiol.  87: 638-646.
         January 2005                                    9-379       DRAFT-DO NOT QUOTE OR CITE

-------
  1       Novak, K.; Skelly, J. M; Schaub, M.; Krauchi, N.; Hug, C.; Landolt, W.; Bleuter, P. (2003) Ozone air pollution and
  2             foliar injury development on native plants of Switzerland. Environ. Pollut. 125: 41-52.
  3       Nunn, A. J.; Reiter, I. M.; Haberle, K.-H.; Werner, H.; Langebartels, C.; Sandermann, H.; Heerdt, C.; Fabian, P.;
  4             Matyssek, R. (2002) "Free-air" ozone canopy fumigation in an old-growth mixed forest: concept and
  5             observations in beech. Phyton 42: 105-119.
  6       Nussbaum, S.; Geissman, M.; Fuhrer, J. (1995) Effects of nitric oxide and ozone on spring wheat (Triticum aestivum).
  1             Water Air Soil Pollut. 85: 1449-1454.
  8       Nussbaum, S.; Geissmann, M.; Fuhrer, J. (1995) Ozone exposure-response relationships for mixtures of perennial
  9             ryegrass and white clover depend on ozone exposure patterns. Atmos. Environ. 29: 989-995.
10       Nussbaum, S.; Ammann, M.; Fuhrer, J.  (1999) Foliar absorption and use of airborne oxidized nitrogen by terrestrial
11             plants.  In: Srivastava, H. S.; Singh, R. P., eds. Nitrogen Nutrition and Plant Growth; pp. 101-172. New Delhi,
12             India: Oxford & ffiH Publishing  Co. Pvt. Ltd.
13       Nussbaum, S.; Bungener, P.; Geissmann, M.; Fuhrer, J. (2000) Plant-plant interactions and soil moisture might be
14             important in determining ozone impacts on grasslands. New Phytol.  147: 327-335.
15       Nussbaum, S.; Geissman, M.; Saurer, M.; Siegwolf, R.; Fuhrer, J. (2000) Ozone and low concentrations of nitric oxide
16             have similar effects on carbon isotope discrimination and gas exchange in leaves of wheat (Triticum aestivum
17             L.). J. Plant Physiol. 156: 741-745.
18       O'Gara, P. J. (1922) Sulfur dioxide and fume problems and their  solution, (presented at the 14th semiannual meeting of
19             the American Institute of Chemical Engineers, Niagara Falls, Canada, June 19-22, 1922).  Presented at: 14th
20             semiannual meeting of the American Institute of Chemical Engineers; June; Niagara Falls, Canada. Proceedings
21             and papers of the meeting summarized in J. Ind. Eng. Chem. 14: 744-745 by J. C. Olsen.
22       Ojanpera'K.; Patsikka, E.; Ylaranta,  T. (1998) Effects of low ozone exposure of spring wheat on net CO2 uptake,
23             Rubisco,  leaf senescence and grain filling. New Phytol. 138: 451-460.
24       Oksanen, E. J. (2001) Increasing tropospheric ozone level reduced birch (Betula pendula) dry mass within a five years
25             period. Water Air Soil Pollut. 130: 947-952.
26       Oksanen, E. (2003a) Physiological responses of birch (Betula pendula) to ozone: a comparison between
27             open-soil-grown trees exposed for six growing seasons and potted seedlings exposed for one season. Tree
28             Physiol. 23:603-614.
29       Oksanen, E. (2003b) Responses of selected birch (Betula pendula Roth) clones to ozone change over time. Plant Cell
30             Environ.  26: 875-886.
31       Oksanen, E.; Holopainen, T. (2001)  Responses of two birch (Betula pendula Roth) clones to different ozone profiles
32             with similar AOT40 exposure. Atmos. Environ. 35: 5245-5254.
33       Oksanen, E.; Saleem, A.  (1999) Ozone exposure results in various carry-over effects and prolonged reduction in
34             biomass in birch (Betula pendula Roth). Plant Cell Environ. 22:  1401-1411.
35       Oksanen, E.; Amores, A.; Kokko, H.; Santamaria, J. M.; Karenlampi, L. (2001a) Genotypic variation in growth and
3 6             physiological responses of Finnish hybrid aspen (Populus tremuloides x P. tremula) to elevated tropospheric
37             ozone concentration. Tree Physiol. 21: 1171-1181.
38       Oksanen, E.; Sober, J.; Karnosky, D. F.  (2001b) Interactions of elevated CO2 and ozone in leaf morphology of aspen
3 9             (Populus tremuloides) and birch  (Betula papyri/era) in aspen FACE experiment. Environ. Pollut. 115: 437-446.
40       Ollerenshaw,  J.  H.; Lyons, T. (1999) Impacts of ozone on the growth and yield of field-grown winter wheat. Environ.
41             Pollut.  106: 67-72.
42       Ollerenshaw,  J.  H.; Lyons, T.; Barnes, J. D. (1999) Impacts of ozone on the growth  and yield of field-grown winter
43             oilseed rape. Environ. Pollut.  104: 53-59.
44       Ollinger, S. V.;  Aber, J. D.; Reich, P. B. (1997) Simulating ozone effects on forest productivity:  interactions among
45             leaf-, canopy-, and stand-level processes. Ecol.  Appl. 7: 1237-1251.
46       Ollinger, S. V.;  Aber, J. D.; Federer, C.  A. (1998) Estimating regional forest productivity and water yield using an
47             ecosystem model linked to a GIS. Landscape Ecol. 13: 323-334.
48       Olszyk, D. M.; Bytnerowicz, A.; Kats, G.; Dawson, P. J.; Wolf, J.; Thompson, C. R. (1986a) Crop effects from air
49             pollutants in air exclusion systems vs. field chambers. J. Environ. Qual. 15: 417-422.
50       Olszyk, D. M.; Kats, G.;  Dawson, P. J.; Bytnerowicz, A.; Wolf, J.; Thompson, C. R. (1986b) Characteristics of air
51             exclusion systems vs. chambers for field air pollution studies. J. Environ.  Qual. 15: 326-334.
52       Olszyk, D. M.; Takemoto, B. K.; Kats, G.; Dawson, P. J.; Morrison, C. L.; Preston,  J. W.; Thompson, C. R. (1992)
53             Effects of open-top chambers on Valencia' orange trees. J. Environ.  Qual. 21: 128-134.
         January 2005                                    9-3 80       DRAFT-DO NOT QUOTE OR CITE

-------
  1       Olszyk, D. M; Tingey, D. T.; Watrad, L.; Seidler, R.; Andersen, C. (2000) Interactive effects of O3 and CO2:
  2             implications for terrestrial ecosystems. In: Singh, S. N., ed. Trace gas emissions and plants. Dordrecht, The
  3             Netherlands: Kluwer Academic Publishers; pp. 97-136.
  4       Olszyk, D. M.; Johnson, M. G.; Phillips, D. L.; Seidler, R. I; Tingey, D. T.; Watrad, L. S. (2001) Interactive effects of
  5             CO2 and O3 on a ponderosa pine/litter/soil mesocosm. Environ. Pollut. 115: 447-462.
  6       Ommen, O. E.; Donnelly, A.; Vanhoutvin, S.; Van Oijen, M.; Manderscheid, R. (1999) Chlorophyll content of spring
  7             wheat flag leaves grown under elevated CO2 concentrations and other environmental stresses within the
  8             'ESPACE-wheat' project. Eur. J. Agron. 10: 197-203.
  9       Ordin, L.; Hall, M. A.; Kindinger, J. I. (1969) Oxidant induced inhibition of enzymes involved in cell wall
10             polysaccharide synthesis. Arch. Environ. Health 18:  623-626.
11       Oren, R.; Ellsworth, D. S.; Johnsen, K. H.; Phillips, N.; Ewers, B. E.; Maier, C.; Schafer, K. V. R.; McCarthy, H.;
12             Hendrey, G.; McNulty, S. G.; Katul, G. G. (2001) Soil fertility limits carbon sequestration by forest ecosystems
13             in a CO2-enriched atmosphere.  Nature (London) 411: 469-472.
14       Orendovici, T.; Skelly, J. M.; Ferdinand, J. A.; Savage, J. E.; Sanz, M.-J.;  Smith, G. C. (2003) Response of native
15             plants of northeastern United States and southern Spain to ozone exposures; determining exposure/response
16             relationships. Environ. Pollut. 125: 31-40.
17       Ormrod, D. P.; Beckerson, D. W. (1986) Polyamines as antiozonants for tomato. HortScience 21: 1070-1071.
18       Ormrod, D. P.; Landry, L.  G.; Conklin, P. L. (1995) Short-term UV-B radiation and ozone exposure effects on
19             aromatic secondary metabolite  accumulation and shoot growth of flavonoid-deficient^rai/Wo/ms- mutants.
20             Physiol. Plant.  93: 602-610.
21       Oshima, R. J. (1975) Development of a system for evaluating and reporting economic crop losses caused by air
22             pollution in California. III. Ozone dosage-crop loss conversion function-alfalfa, sweet corn. IIIA. Procedures
23             for production, ozone effects on alfalfa and sweet corn, and evaluation of these systems. Sacramento, CA:
24             California Air Resources Board; report no. ARB-R-3-690-75-35. Available from: NTIS, Springfield, VA;
25             PB-272994.
26       Oshima, R. J.; Poe, M. P.; Braegelmann, P. K.; Baldwin, D. W.; Van Way, V (1976) Ozone dosage-crop loss function
27             for alfalfa: a standardized method for assessing crop losses from air pollutants. J. Air Pollut. Control Assoc. 26:
28             861-865.
29       Oshima, R. J.; Braegelmann, P. K.; Baldwin, D. W.; Van Way, V; Taylor, O. C. (1977) Responses of five cultivars of
30             fresh market tomato to ozone: a contrast of cultivar screening with foliar injury and yield. J.  Am. Soc. Hortic.
31             Sci. 102: 286-289.
32       Paakkonen, E.; Holopainen, T. (1995) Influence of nitrogen supply on the response of clones of birch (Betula pendula
33             Roth) to ozone. New Phytol. 129: 595-603.
34       Paakkonen, E.; Holopainen, T.; Karenlampi, L. (1996) Relationships between open-field ozone exposures and growth
35             and senescence of birch (Betula pendula and Betula pubescens). In: Karenlampi, L.; Skarby, L., eds. Critical
36             levels for ozone in Europe: testing and finalizing the concepts: UN-ECE workshop report. UN-ECE Convention
37             on Long-Range Transboundary Air Pollution; April; Kuopio, Finland. Kuopio, Finland: University of Kuopio,
38             Department of Ecology and Environmental Science;  pp. 298-302.
39       Paakkonen, E.; Gunthardt-Georg, M. S.; Holopainen, T. (1998) Responses of leaf processes in a sensitive birch (Betula
40             pendula Roth) clone to ozone combined with drought. Ann. Bot. 82: 49-59.
41       Paludan-Muller, G.; Saxe,  H.; Leverenz, J. W. (1999) Responses to ozone in 12 provenances of European beech
42             (Fagus sylvatica): genotypic variation and chamber effects on photosynthesis and dry-matter partitioning. New
43             Phytol. 144: 261-273.
44       Panek, J. A.; Goldstein, A. H. (2001) Responses of stomatal conductance to drought in ponderosa pine: implications
45             for carbon and ozone uptake. Tree Physiol. 21: 337-344.
46       Panek, J.; Kurpius, M. R.;  Goldstein, A. H. (2002) An evaluation of ozone exposure metrics for a seasonally
47             drought-stressed ponderosa pine ecosystem. Environ. Pollut. 117: 93-100.
48       Panek, J. A.; Baldocchi, D. D.; Goldstein, A. H. (2003) The need for spatially and functionally integrated models of
49             ozone deposition to Sierra Nevada forests. In: Bytnerowicz, A.; Arbaugh, M. J.; Alonso, R., eds. Ozone Air
50             Pollution in the Sierra Nevada: Distribution and Effects on Forests. New York, NY:  Elsevier; pp. 325-357.
51             (Developments in Environmental Science, v. 2).
52       Patterson, M. T.; Rundel, P. W. (1995) Stand characteristics of ozone-stressed populations of Pinus jeffreyi (Pinaceae):
53             extent, development, and physiological consequences of visible injury. Am. J. Bot. 82: 150-158.
         January 2005                                    9-3 81        DRAFT-DO NOT QUOTE OR CITE

-------
  1       Pauls, K. P.; Thompson, J. E. (1980) In vitro simulation of senescence-related membrane damage by ozone-induced
  2             lipid peroxidation. Nature (London) 283: 504-506.
  3       Pearce, D. W. (1993) Economic values and the natural world. London, United Kingdom: Earthscan Publications, Ltd.
  4       Pearson, M; Mansfield, T. A. (1993) Interacting effects of ozone and water stress on the stomatal resistance of beech
  5             (Fagus sylvatica L.). New Phytol. 123: 351-358.
  6       Pearson, M.; Mansfield, T. A. (1994) Effects of exposure to ozone and water stress on the following season's growth of
  7             beech (Fagus sylvatica L.). New Phytol. 126: 511-515.
  8       Pearson, S.; Davison, A. W.; Reiling, K.; Ashenden, T.; Ollerenshaw, J. H. (1996) The effects of different ozone
  9             exposures on three  contrasting populations ofPlantago major. New Phytol. 132: 493-502.
10       Pel, Z.-M.; Murata, Y.; Benning, G.; Thomine, S.; Kltisener, B.; Allen, G. J.; Grill, E.; Schroeder, J. I. (2000) Calcium
11             channels activated by hydrogen peroxide mediate abscisic acid signalling in guard cells. Nature (London) 406:
12             731-734.
13       Pell, E. J.; Pearson, N. S. (1984) Ozone-induced reduction in quantity and quality of two potato cultivars. Environ.
14             Pollut. Ser. A 35: 345-352.
15       Pell, E. J.; Eckardt, N.; Enyedi, A. J. (1992) Timing of ozone stress and resulting status of ribulose bisphosphate
16             carboxylase/oxygenase and associated net photosynthesis. New Phytol. 120: 397-405.
17       Pell, E. J.; Eckardt, N. A.; Click, R. E. (1994) Biochemical and molecular basis for impairment of photosynthetic
18             potential. Photosynth. Res. 39: 453-462.
19       Pell, E. J.; Sinn, J. P.; Johansen, C. V. (1995) Nitrogen supply as a limiting factor determining the sensitivity ofPopulus
20             tremuloides Michx. to ozone stress.  New Phytol. 130: 437-446.
21       Pell, E. J.; Schlagnhaufer, C. D.; Arteca, R. N. (1997) Ozone-induced oxidative stress: mechanisms of action and
22             reaction. Physiol. Plant. 100: 264-273.
23       Pell, E. J.; Sinn, J. P.; Brendley, B. W.; Samuelson, L.; Vinten-Johansen, C.; Tien, M.; Skillman, J. (1999) Differential
24             response of four tree species to ozone-induced acceleration of foliar senescence. Plant Cell Environ. 22: 779-790.
25       Pellinen, R.; Palva, T.; Kangasjavi, J. (1999) Subcellular localization of ozone-induced hydrogen peroxide production in
26             birch (Betula pendula) leaf cells. Plant. J. 20: 349-356.
27       Pelloux, J.; Jolivet, Y.; Fontaine, V.; Banvoy, J.; Dizengremel, P. (2001)  Changes inRubisco and Rubisco activase gene
28             expression and polypeptide content in Pinus halepensis M. subjected to ozone and drought. Plant Cell Environ.
29             24: 123-131.
30       Penel, C. (1986) The role  of calcium in the control of peroxidase activity. In: Greppin, H.; Penel, C.; Gasper, T., eds.
31             Molecular and physiological aspects of plant peroxidases. Geneva, Switzerland: Universite de Genieve, Centre
32             de Botanique; pp. 155-164.
33       Percy, K. E.; Awmack, C. S.; Lindroth, R. L.; Kubiske, M. E.; Kopper, B. J.; Isebrands, J. G.; Pregitzer, K. S.; Hendry,
34             G. R.; Dickson, R.  E.; Zak, D. R.; Oksanen, E.; Sober, J.; Harrington, R.; Karnosky, D. F. (2002) Altered
35             performance of forest pests under atmospheres enriched with CO2  and O3. Nature (London) 420: 403-407.
36       Percy, K. E.; Karnosky, D. F.; Mankovska, B. (2003) Interactive O3 and CO2 effects on leaf surface physicochemical
37             characteristics in three aspen clones: two-year results from the Aspen FACE experiment. In: Karnosky, D. F.;
38             Percy, K. E.; Chappelka, A. H.; Simpson, C.  J.; Pikkarainen, J., eds. Air pollution, global  change and forests in
39             the new millenium. New York, NY: Elsevier Science. (Developments in Environmental Science, 3).
40       Persson, K.; Danielsson, H.; Selden, G.; Pleijel, H. (2003) The effects of tropospheric ozone and elevated carbon
41             dioxide on potato (Solanum tuberosum L. cv. Bintje) growth and yield. Sci. Total Environ. 310:  191-201.
42       Peters, J.  L.; Castillo, F. J.; Heath, R. L. (1988) Alteration of extracellular enzymes in pinto bean leaves upon exposure
43             to air pollutants, ozone and sulfur dioxide. Plant Physiol.  89: 159-164.
44       Peterson, D. L.; Arbaugh,  M. J.; Wakefield, V. A.; Miller, P. R.  (1987) Evidence of growth reduction in ozone-injured
45             Jeffrey pine (Pinus jeffreyi Grev. and Balf.) in Sequoia and Kings  Canyon National Parks. JAPCA 37: 906-912.
46       Peterson, D. L.; Arbaugh,  M. J.; Robinson, L. J. (1991) Regional growth changes in ozone-stressed ponderosa pine
47             (Pinus ponderosa)  in the Sierra Nevada, California, USA. Holocene 1: 50-61.
48       Peterson, D. L.; Silsbee, D. G.; Poth, M.; Arbaugh, M. J.; Biles, F. E. (1995) Growth response of bigcone Douglas fir
49             (Pseudotsuga macrocarpa) to long-term ozone exposure in southern California. J. Air Waste Manage. Assoc. 45:
50             36-45.
51       Pfirrmann, T.; Barnes, J. D. (1993) Interactive effects of CO2 and O3 on growth of radish (Raphanus sativusL.).
52             Vegetatio 104-105: 464-466.
53       Phillips, R. L.; Zak, D. R.; Holmes, W. E.; White, D. C. (2002) Microbial community composition and function beneath
54             temperate trees exposed to elevated atmospheric carbon dioxide and ozone. Oecologia 131: 236-244.


         January 2005                                   9-3 82       DRAFT-DO NOT QUOTE OR CITE

-------
  1       Pimm, S. L. (1997) The value of everything. Nature (London) 387: 231-232.
  2       Pino, M. E.; Mudd, J. B.; Bailey-Serres, J. (1995) Ozone-induced alterations in the accumulation of newly synthesized
  3             proteins in leaves of maize. Plant Physiol. 108: 777-785.
  4       Pitcher, L. H.; Zilinskas, B. A. (1996) Overexpression of copper/zinc superoxide dismutase in the cytosol of transgenic
  5             tobacco confers partial resistance to ozone-induced foliar necrosis. Plant Physiol. 110: 583-588.
  6       Pitcher, L. H.; Brennan, E.; Zilinskas, B. A. (1992) The antiozonant ethylenediurea does not act via superoxide
  7       dismutase induction in bean.  Plant Physiol. 99: 1388-1392.
  8       Pitelka, L. F. (1988) Evolutionary responses of plants to anthropogenic pollutants. Trends in Ecology and Evolution 3:
  9             233-236.
10       Pleijel, H. (1998) A suggestion of a simple transfer function for the use of ozone monitoring data in dose-response
11             relationships obtained using open-top chambers.  Water Air Soil Pollut. 102: 61-74.
12       Pleijel, H.; Danielsson, H.  (1997) Growth of 27 herbs and grasses in relation to ozone exposure and plant strategy. New
13             Phytol. 135: 361-367.
14       Pleijel, H.; Skarby, L.; Ojanpera, K.; Sellden, G. (1994) Exposure of oats, avena sativa L., to filtered and unfiltered air
15             in open-top chambers: Effects on grain yield and quality. Environ. Pollut. 86:  129-134.
16       Pleijel, H.; Wallin, G.; Karlsson, P.  E.; Skarby, L.; Sellden, G. (1994) Ozone deposition to an oat crop (Avena sativa L.)
17             grown in open-top chambers and in the ambient air, Atmos. Environ. 28:  1971-1979.
18       Pleijel, H.; Wallin, G.; Karlsson, P.; Skarby, L.; Sellden, G. (1995) Gradients of ozone at a forest site and over a field
19             crop - consequences for the AOT4O concept of critical level. Water Air Soil Pollut. 85: 203 3 -203 8.
20       Pleijel, H.; Karlsson, G. P.; Danielsson, H.; Sellden, G.  (1995) Surface wetness enhances ozone deposition to a pasture
21             canopy. Atmos. Environ. 29: 3391-3393.
22       Pleijel, H.; Karlsson, G. P.; Sild, E.; Danielsson, H.; Skarby, L.; Sellden, G. (1996) Exposure of a grass-clover mixture
23             to ozone in open-top chambers-effects on yield, quality and botanical composition Agric. Ecosyst. Environ. 59:
24             55-62.
25       Pleijel, H.; Ojanpera, K.; Mortensen, L. (1997) Effects of tropospheric ozone on the yield and grain protein content of
26             spring wheat (Triticum aestivum L.) in the nordic countries. Acta Agric. Scand. Sect. B 47: 20-25.
27       Pleijel, H.; Danielsson, H.; Gelang,  J.; Sild, E.; Sellden, G. (1998) Growth stage dependence of the grain yield response
28             to ozone in spring wheat (Triticum aestivum L.).  Agric. Ecosyst. Environ. 70:  61-68.
29       Pleijel, H.; Mortensen, L.;  Fuhrer, J.; Ojanpera, K.; Danielsson, H. (1999) Grain protein accumulation in relation to
3 0             grain yield of spring wheat (Triticum aestivum L.) grown in open-top chambers with different concentrations of
31             ozone, carbon dioxide and water availability. Agric. Ecosyst. Environ. 72: 265-270.
32       Pleijel, H.; Norberg, P. A.; Sellden, G.; Skarby, L. (1999) Tropospheric ozone decreases biomass production in radish
33             plants (Raphanus sativus) grown in rural south-west Sweden. Environ. Pollut. 106: 143-147.
34       Pleijel, H.; Danielsson, H.; Karlsson, G. P.; Gelang, J.; Karlsson, P.  E.; Sellden, G. (2000) An ozone flux-response
35             relationship for wheat. Environ. Pollut. 109: 453-462.
36       Pleijel, H.; Gelang, J.; Sild, E.; Danielsson, H.; Younis, S.; Karlsson, P.-E.; Wallin, G.; Skarby, L.; Sellden, G. (2000)
3 7             Effects of elevated carbon dioxide, ozone and water availability on spring wheat growth and yield. Physiol. Plant.
38             108:61-70.
39       Pleijel, H.; Danielsson, H.; Vandermeiren, K.; Blum, C.; Colls, J.; Ojanpera, K. (2002) Stomatal conductance and ozone
40             exposure in relation to potato tuber yield-results from the European CHIP programme. Eur. J. Agron. 17:
41             303-317.
42       Plochl, M.; Lyons, T.; Ollerenshaw, J.; Barnes, J. (2000) Simulating ozone detoxification in the leaf apoplast through
43             the direct reaction with ascorbate. Planta 210: 454-467.
44       Polasky, S.,  ed. (2001) The economics of biodiversity conservation. Burlington, VT: Ashgate Publishers.
45       Polle, A.; Pell, E. J. (1999) Role of carbon dioxide in modifying the plant response to ozone. In: Luo, Y.; Mooney, H.
46             A., eds. Carbon dioxide and environmental stress. San Diego, CA: Academic Press; pp. 193-213.
47       Polle, A.; Rennenberg, H. (1991) Superoxide dismutase activity in needles of Scots pine and Norway spruce under field
48             and changer conditions: lack of ozone effects. New Phytol. 117:  335-343.
49       Polle, A.; Krings, B.; Rennenberg, H. (1989) Superoxide dismutase activity in needles of Norwegian spruce trees (Picea
50             abies 1.) Plant Physiol. 90: 1310-1315.
51       Poorter, H.;  Perez-Soba, M. (2001)  The growth response of plants to elevated CO2 under non-optimal environmental
52             conditions. Oecologia 129: 1-20.
         January 2005                                     9-3 83        DRAFT-DO NOT QUOTE OR CITE

-------
  1       Posthumus, A. C. (1998) Air pollution and global change: significance and prospectives. In: De Kok, L. I; Stulen, I.,
  2             eds. Responses of plant metabolism to air pollution and global change. Leiden, The Netherlands: Backhuys
  3             Publishers; pp. 3-14.
  4       Poth, M. A.; Fenn, M. E. (1998) Mature ponderosa pine nutrient use and allocation responses to air pollution. In:
  5             Bytnerowicz, A.; Arbaugh, M. I; Schilling, S. L. Proceedings of the International Symposium on Air Pollution
  6             and Climate Change Effects on Forest Ecosystems; February, 1996; Riverside, CA. Albany, CA: USDA Forest
  7             Service, Pacific Southwest General Technical Report no. PSW-GTR 166; pp. 239-247.
  8       Potter, L.; Foot, J. P.; Caporn, S. J. M.; Lee, J. A. (1996) The effects of long-term elevated ozone concentrations on the
  9             growth and photosynthesis of Sphagnum recurvum and Polytrichum commune. New Phytol. 134: 649-656.
10       Powell, M. C.; Muntifering, R. B.; Lin, J. C.; Chappelka, A. H. (2003) Yield and nutritive quality of sericea lespedeza
11             (Lespedeza cuneata) and little bluestem (Schizachyrium scoparium) exposured to ground-level ozone. Environ.
12             Pollut. 122:313-322.
13       Pronos, J.; Merrill, L.; Dahlsten, D. (1999) Insects and pathogens in a pollution-stressed forest. In: Miller, P. R.;
14             McBride, J. R., eds. Oxidant air pollution impacts in the montane forests of southern California. New York, NY:
15             Springer, pp. 317-337.
16       Prozherina, N.; Freiwald, V.; Rousi, M.; Oksanen, E. (2003) Interactive effect of springtime frost and elevated ozone on
17             early growth, foliar injuries and leaf structure of birch (Betulapendula). New Phytol. 159: 623-636.
18       Pryor, W. A. (1992) How far does ozone penetrate into the pulmonary air/tissue boundary before it reacts? Free Radic.
19             Biol. Med. 12: 83-88.
20       Qui, Z.; Chappelka, A. H.; Somers, G. L.; Lockaby, B. G.; Meldahl, R. S. (1993) Effects of ozone and simulated acidic
21             precipitation on ectomycorrhizal formation on loblolly pine seedlings. Environ. Exp. Bot. 33: 423-431.
22       Rabinowitch, E. I. (1945) Photosynthesis and related processes. Vol. 1: chemistry of photosynthesis, chemosynthesis
23             and related processes in vitro and in vivo. New York, NY: Interscience Publishing, Inc.
24       Rafarel, C. R.; Ashenden, T. W. (1991) A facility for the large-scale exposure of plants to gaseous atmospheric
25             pollutants. New Phytol. 117: 345-349.
26       Randall, A. (1991) The value of biodiversity. Ambio 20: 64-68.
27       Rao, M.; Davis, K. R. (1999) Ozone-induced cell death occurs via two distinct mechanisms mArabidopsis'. the role of
28             salicylic acid. Plant J. 17: 603-614.
29       Rao, M. V.; Hale, B. A.; Ormrod, D. P. (1995) Amelioration of ozone-induced oxidative damage in wheat plants grown
30             under high carbon dioxide. Role of antioxidant enzymes. Plant Physiol. 109: 421-432.
31       Rawlings, J. O.; Lesser, V. M.; Heagle, A. S.; Heck, W. W. (1988) Alternative ozone dose metrics to characterize ozone
32             impact on crop yield loss. J. Environ. Qual. 17: 285-291.
33       Read, J. J.; Morgan, J. A. (1996) Growth and partitioning mPascopyrum smithii (C3) and Bouteloua gracilis (C4) as
34             influenced by carbon dioxide and temperature. Ann. Bot. (London, UK) 77: 487-496.
35       Reams, G. A.; VanDeusen, P. C.; Lucier, A. A. (1995) Ambient ozone and loblolly pines. Nature (London) 378:
36             449-450.
37       Rebbeck, J. (1996) Chronic ozone effects on three northeastern hardwood species: growth and biomass. Can.  J. For.
38             Res. 26: 1788-1798.
39       Rebbeck, J.; Jensen, K. F.; Greenwood, M. S. (1993) Ozone effects on grafted mature and juvenile red spruce:
40             photosynthesis, stomatal conductance, and chlorophyll concentration.  Can. J. For. Res. 23: 450-456.
41       Reddy, G. N.; Arteca, R. N.; Dai, Y.-R.; Flores, H. E.; Negm, F. B.; Pell, E. J. (1993) Changes in ethylene and
42             polyamines in relation to mRNA levels of the large and small subunits of ribulose bisphosphate
43             carboxylase/oxygenase in ozone-stressed potato foliage. Plant Cell Environ. 16: 819-826.
44       Reich, P. B. (1983) Effects of low concentrations of O3 on net photosynthesis, dark respiration, and chlorophyll contents
45             in aging hybrid poplar leaves. Plant Physiol. 73: 291-296.
46       Reich, P. B.; Lassoie, J. P. (1984) Effects of low level O3 exposure on leaf diffusive conductance and water-use
47             efficiency in hybrid poplar. Plant Cell Environ. 7: 661-668.
48       Reichenauer, T. G.; Goodman, B. A.; Kostecki, P.; Soja, G. (1998) Ozone sensitivity in Triticum durum and T. aestivum
49             with respect to leaf injury, photosynthetic activity and free radical content. Physiol. Plant. 104: 681-686.
50       Reid, C. D.; Fiscus, E. L. (1998) Effects of elevated [CO2] and/or ozone on limitations to CO2 assimilation in soybean
51             (Glycine max). J. Exp. Bot. 49: 885-895.
52       Reid, C. D.; Fiscus, E. L.; Burkey, K. O. (1998) Combined effects of chronic ozone and elevated CO2 on rubisco
53             activity and leaf components in soybean (Glycine max). J. Exp. Bot. 49: 1999-2011.
         January 2005                                     9-3 84       DRAFT-DO NOT QUOTE OR CITE

-------
  1       Railing, K.; Davison, A. W. (1992a) Spatial variation in ozone resistance of British populations ofPlantago major L.
  2             New Phytol. 122: 699-708.
  3       Reiling, K.; Davison, A. W. (1992b) Effects of a short ozone exposure given at different stages in the development of
  4             Plantago major L. New Phytol. 121: 643-647.
  5       Reiling, K.; Davison, A. W. (1992c) The response of native, herbaceous species to ozone: growth and fluorescence
  6             screening. New Phytol. 120: 29-37.
  7       Reiling, K.; Davison, A. W. (1994) Effects of exposure to ozone at different stages in the development ofPlantago
  8             major L. on chlorophyll fluorescence and gas exchange. New Phytol. 128:  509-514.
  9       Reimer, I; Whittaker,  J. B. (1989) Air pollution and insect herbivores: observed interactions and possible mechanisms.
10             In: Bernays, E. A., ed.  Insect-plant interactions, v. 1. Boca Raton, FL: CRC Press, pp. 73-105.
11       Reiner, S.; Wiltshire, J. J. J.; Wright, C. J.; Colls, J. J. (1996) The impact of ozone and drought on the water relations of
12             ash trees (Fraxinus excelsior L.). J. Plant Physiol. 148: 166-171.
13       Reinert, R. A.; Ho, M. C. (1995) Vegetative growth of soybean as affected by elevated carbon dioxide and ozone.
14             Environ. Pollut. 89: 89-96.
15       Reinert, R. A.; Eason,  G.; Barton, J. (1997) Growth and fruiting of tomato as influenced by elevated carbon dioxide and
16             ozone. New Phytol. 137: 411-420.
17       Renaud, J. P.; Allard, G.; Mauffette, Y. (1997) Effects of ozone on yield, growth, and root starch concentrations of two
18             alfalfa (Medicago sativa L.)  cultivars. Environ. Pollut. 95: 273-281.
19       Retzlaff, W. A.; Weinstein, D. A.; Laurence, J. A.; Gollands, B. (1996) Simulated root dynamics of a 160-year-old sugar
20             maple (Acer saccharum Marsh.) tree with and without ozone exposure using the TREGRO model. Tree Physiol.
21             16:915-921.
22       Retzlaff, W. A.; Williams, L.  E; DeJong, T. M. (1997) Growth and yield response of commercial bearing-age
23             'Casselman' plum trees to various ozone partial pressures. J. Environ. Qual. 26: 858-865.
24       Retzlaff, W. A.; Arthur, M. A.; Grulke, N. E.; Weinstein, D. A.; Gollands, B. (2000) Use of a single-tree simulation
25             model to predict effects of ozone and drought on growth of a white fir tree. Tree Physiol. 20: 195-202.
26       Ribas, A.; Penuelas, J. (2000) Effects of ethylene diurea as a protective antiozonant on beans (Phaseolus vulgaris cv lit)
27             exposed to different tropospheric ozone doses in Catalonia (NE Spain). Water Air Soil Pollut. 117: 263-271.
28       Rich, S.; Ames, R.; Zukel, J. W. (1974) 1,4-oxanthin derivatives protect plants against ozone. Plant Dis. Rep.  58:
29             162-164.
3 0       Robinson, J. M.; Britz, S. J. (2000) Tolerance of a field grown soybean cultivar to elevated ozone level is concurrent with
31             higher leaflet ascorbic acid level, higher ascorbate-dehydroascorbate redox status, and long term photosynthetic
32             productivity. Photosynth. Res. 64: 77-87.
33       Robinson, M. F.; Heath, J.; Mansfield, T. A. (1998) Disturbance in stomatal behaviour caused by air pollutants. J. Exp.
34             Bot. 49: 461-469.
35       Roitsch, T. (1999) Source-sink regulation by sugar and stress. Curr. Opin. Plant Biol. 2:  198-206.
36       Rondon, A.; Johansson, C.; Granat, L. (1993) Dry deposition of nitrogen-dioxide  and ozone to coniferous forests. J.
37             Geophys. Res. [Atmos.] 98(D3): 5159-5172.
38       Rosen, P. M.; Runeckles, V. C. (1976) Interaction of ozone and greenhouse whitefly in plant injury. Environ.  Conserv. 3:
39             70-71.
40       Rotter, R.; Van De Geijn, S. C. (1999) Climate change effects on plant growth, crop yield and livestock. Clim. Change
41             43:651-681.
42       Rowland-Bamford, A. J. (2000) Plant responses to changing carbon dioxide and temperature. In:  Singh, S. N., ed. Trace
43             gas emissions and plants. Dordrecht, The Netherlands: Kluwer Academic Publishers; pp. 63-74.
44       Rudorff, B. F. T.; Mulchi, C. L.; Daughtry, C. S. T.; Lee, E. H. (1996) Growth, radiation use efficiency, and canopy
45             reflectance of wheat and corn grown under elevated ozone and carbon dioxide atmospheres. Remote Sens.
46             Environ. 55: 163-173.
47       Rudorff, B. F. T.; Mulchi, C. L.; Fenny, P.; Lee, E. H.; Rowland, R. (1996) Wheat grain quality under enhanced
48             tropospheric CO2 and O3 concentrations. J. Environ. Qual. 25: 1384-1388.
49       Rudorff, B. F. T.; Mulchi, C. L.; Lee, E. H.; Rowland, R.; Pausch, R. (1996) Effects of enhanced O3 and CO2  enrichment
50             on plant characteristics in wheat and corn. Environ. Pollut. 94: 53-60.
51       Rudorff, B. F. T.; Mulchi, C. L.; Lee, E. H. (2000) Plant responses to elevated CO2 and interactions with O3. In:  Singh, S.
52             N., ed. Trace gas emissions and plants. Dordrecht, The Netherlands: Kluwer Academic Publishers; pp.  155-179.
53       Runeckles, V. C. (1974) Dosage of air pollutants and damage to vegetation. Environ. Conserv.  1: 305-308.
         January 2005                                     9-385       DRAFT-DO NOT QUOTE OR CITE

-------
  1       Runeckles, V. C. (2002) Air pollution and climate change. In: Bell, J. N. B.; Treshow, M, eds. Air Pollution and Plant
  2             Life. 2nd ed. Chichester, United Kingdom: John Wiley & Sons Ltd.; pp. 431-454.
  3       Runeckles, V. C.; Bowen, P. A. (2000) The use of calibrated passive monitors to assess crop loss due to ozone in rural
  4             locations. In: Agrawal, S. B.; Agrawal, M., eds. Environmental pollution and plant responses. Boca Raton, FL:
  5             Lewis Publishers; pp. 307-316.
  6       Runeckles, V. C.; Krupa, S. V. (1994) The impact of UV-B radiation and ozone on terrestrial vegetation. Environ. Pollut.
  7             83: 191-213.
  8       Runeckles, V. C.; Palmer, K. (1987) Pretreatment with nitrogen dioxide modifies plant response to ozone. Atmos.
  9             Environ. 21:717-719.
10       Runeckles, V. C.; Rosen, P. M. (1974) Effects of pretreatment with low ozone concentrations on ozone injury to bean
11             and mint. Can. J. Bot. 52: 2607-2610.
12       Runeckles, V. C.; Vaartnou, M. (1997) EPR evidence for superoxide anion formation in leaves during exposure to low
13             levels of ozone. Plant Cell Environ. 20: 306-314.
14       Runeckles, V. C.; Wright, E. F. (1996) Delayed impact of chronic ozone stress on young Douglas-fir grown under field
15             conditions. Can. J. For. Res. 26: 629-638.
16       Runeckles, V. C.; Wright, E. F.; White, D. (1990) A chamberless field exposure system for determining the effects of
17             gaseous air pollutants on crop growth and yield. Environ. Pollut. 63: 61 -77.
18       Rusch, H.; Laurence, J. A. (1993) Interactive effects of ozone and powdery mildew on pea seedlings. Phytopathology 83:
19             1258-1263.
20       Sah, D. N.; Von Tiedemann, A.; Fehrmann, H. (1993)  Effects of ozone exposure on resistance of wheat genotypes to
21             Phrenophora tritici-repentis. J. Phytopathol. 138:  249-256.
22       Sakaki, T.; Kondo, N.; Sugahara, K. (1983) Breakdown of photosynthetic pigments and lipids in spinach leaves with
23             ozone fumigation: role of active oxygens. Physiol. Plant. 59: 28-34.
24       Sakaki, T.; Ohnishi, J.-I; Kondo, N.; Yamada, M. (1985) Polar and neutral lipid changes in spinach leaves with ozone
25             fumigation: triacylglyerol synthesis from polar lipids. Plant Cell Physiol. 26: 253-262.
26       Sakaki, T.; Saito, K.; Kawaguchi, A.; Kondo, N.; Yamada, M. (1990) Conversion of monogalactosylidacyglyerol to
27             triacylglycerols in ozone-fumigated spinach leaves. Plant Physiol. 94: 766-772.
28       Salardino, D. H.; Carroll, J. J. (1998)  Correlation between ozone exposure and visible foliar injury in ponderosa  and
29             Jeffrey pines. Atmos. Environ. 32: 3001-3010.
30       Saleem, A.; Loponen, J.; Pihlaja, K.; Oksanen, E. (2001) Effects of long-term open-field ozone exposure on leaf
31             phenolics of European silver birch (Betula pendula Roth). J. Chem. Ecol. 27: 1049-1062.
32       Salisbury, F. B.; Ross, C. (1969) Plant physiology. Belmont, CA: Wadsworth Publishing Company, Inc.
33       Samuelson, L. J. (1994) Ozone-exposure responses of black cherry and red maple seedlings. Environ. Exp. Bot. 34:
34             355-362.
35       Samuelson, L. (2001) Relating ozone uptake in forest trees to risk assessment.  Human Ecol. Risk Assess. 7: 1161-1167.
36       Samuelson, L. J.; Edwards, G.  S. (1993) A comparison of sensitivity to ozone in seedlings and trees of Quercus rubra L.
37             New Phytol. 125: 373-379.
38       Samuelson, L. J.; Kelly, J. M. (1996)  Carbon partitioning and allocation in northern red oak seedlings and mature trees in
39             response to ozone. Tree Physiol. 16: 853-858.
40       Samuelson, L. J.; Kelly, J. M. (1997)  Ozone uptake in Prunus semtina, Acer rubrum and Quercus rubra forest trees of
41             different sizes. New Phytol.  136: 255-264.
42       Samuelson, L.; Kelly, J. M. (2001) Scaling ozone effects from seedlings to forest trees. Tansley review no. 21. New
43             Phytol. 149: 21-41.
44       Samuelson, L. J.; Kelly, J. M.; Mays,  P. A.; Edwards, G.  S. (1996) Growth and nutrition of Quercus rubra L. seedlings
45             and mature trees after three seasons of ozone exposure. Environ. Pollut. 91: 317-323.
46       Sandermann, H., Jr. (1996) Ozone and plant health. Annu. Rev. Phytopathol. 34: 347-366.
47       Sandermann, H., Jr. (1998) Ozone: an air pollutant acting as a plant-signaling molecule. Naturwissenschaften 85:
48             369-375.
49       Sandermann, H., Jr. (2000) Ozone/biotic disease interactions: molecular biomarkers as a new experimental tool. Environ.
50             Pollut. 108: 327-332.
51       Sanders, G. E.; Skarby, L.; Ashmore,  M. R.; Fuhrer, J. (1995) Establishing critical levels to the effects of air pollution on
52             vegetation. Water Air Soil Pollut. 85: 189-200.
53       Saxe, H.; Ellsworth, D. S.; Heath, J. (1998) Tree and forest functioning in an enriched CO2 atmosphere. New Phytol.
54             139: 395-436.


         January 2005                                     9-3 86      DRAFT-DO NOT QUOTE OR CITE

-------
  1       Sayre, R. G.; Fahey, T. J. (1999) Effects of rainfall acidity and ozone on foliar leaching in red spruce Picearubens. Can.
  2             J. For. Res. 29: 486-496.
  3       Scagel, C. F.; Andersen, C. P. (1997) Seasonal changes in root and soil respiration of ozone-exposed ponderosa pine
  4             (Pinusponderosa) grown in different substrates. New Phytol. 136: 627-643.
  5       Scandalios, J. G. (1993) Oxygen stress and superoxide dismutases. Plant Physiol. 101:  7-12.
  6       Schaub, M; Skelly, J. M; Steiner, K. C.; Davis, D. D.; Pennypacker, S. P.; Zhang, J.; Ferdinand, J. A.; Savage, J. E.;
  7             Stevenson, R. E. (2003) Physiological and foliar injury responses ofPmnus serotina, Fraxinus americana, and
  8             Acer rubrum seedlings to varying soil moisture and ozone. Environ. Pollut. 124: 307-320.
  9       Scherzer, A. J.; Rebbeck, J.; Boerner, R. E. J. (1998) Folia nitrogen dynamics and decomposition of yellow-poplar and
10             eastern white pine during four seasons of exposure to elevated ozone and carbon dioxide. For. Ecol. Manage. 109:
11             355-366.
12       Schmieden, U.; Wild, A. (1995) The contribution of ozone to forest decline. Physiol. Plant. 94: 371-378.
13       Schnitzler, J.-P.; Langebartels, C.; Heller, W.; Liu, J.; Lippert, M.; Dohring, T.; Bahnweg, G.; Sandermann, H. (1999)
14             Ameliorating effect of UV-B radiation on the response of Norway spruce and Scots pine to ambient ozone
15             concentrations. Global Change Biol. 5: 83-94.
16       Schnug, E. (1998) Response of plant metabolism to air pollution and global change - impact on agriculture. In: De Kok,
17             L.  J.; Stolen, I., eds. Responses of plant metabolism to air pollution and global change. Leiden, The Netherlands:
18             Backhuys Publishers; pp. 15-22.
19       Schopfer, P. (1994) Histochemical demonstration and localization of H2O2 in organs of higher plants by tissue printing
20             on nitrocellulose paper. Plant Physiol. 104: 1269-1275.
21       Schraudner, M.; Ernst, D.; Langebartels, C.; Sandermann, H., Jr. (1992) Biochemical plant responses to ozone. III.
22             Activation of the defense-related proteins p-l,3-glucanase and chitinase in tobacco leaves. Plant Physiol. 99:
23             1321-1328.
24       Schraudner, M.; Langebartels, C.; Sandermann, H., Jr. (1996) Plant defence systems and ozone. Biochem. Soc. Trans.
25             24:456-461.
26       Schulte-Hosted, S.; Darrall, N. M.; Blank, L. W.; Wellburn, A. R. (1988) Air pollution and plant metabolism. New York,
27             NY: Elsivier Applied Science.
28       Segerson, K. (1987) Impacts of ozone and acid rain: discussion. Am. J. Agric. Econ. 69: 970-971.
29       Segerson, K. (1991) Air pollution and agriculture: a review and evaluation of policy interactions. In: Just, R. E.;
30             Bockstael, N., eds. Commodity and resource policies in agricultural systems. Berlin, Germany: Springer-Verlag;
31             pp. 349-367.
32       Seufert, G.; Arndt, U. (1985) Open-top-Kammern als teil eines konzepts zur okosystemaren untersuchung der neuartigen
33             waldschaden. Allg. Forstz. 40: 13-18.
34       Shan, Y.; Feng, Z.; Izuta, T.; Aoki, M.; Totsuka, T. (1996) The individual and combined effects of ozone and simulated
35             acid rain on growth, gas exchange and water-use efficiency of Pinus armandi Franch. Environ. Pollut. 91:
36             355-361.
37       Sharma, P.; Sober,  A.; Sober, J.; Podila,  G. K.; Kubiske, M. E.; Mattson, W. J.; Isebrands, J. G.; Karnosky, D. F. (2003)
3 8             Moderation of [CO2]-induced gas exchange responses by elevated tropospheric  O3 in trembling aspen and sugar
39             maple. Ekologia (Journal of Ecology)  (Bratislava): in press.
40       Shaw, M. R.; Zavaleta, E. S.; Chiariello, N. R.; Cleland, E. E.; Mooney, H.  A.; Field, C. B. (2002) Grassland responses
41             to  global environmental changes suppressed by elevated CO2. Science (Washington, DC, U.S.) 298: 1987-1990.
42       Sheng, Y.; Podila, G. K.; Karnosky, D. F. (1997) Differences in O3-induced superoxide dismutase and glutathione
43             antioxidant expression in O3 tolerant and sensitive trembling aspen (Populus tremuloides Michx.) clones. For.
44             Gen. 4:  25-33.
45       Sheppard, L. J.; Cape, J. N. (1999) Forest growth responses to the pollution climate of the 21st century: proceedings of
46             the 18th IUFRO workshop on air pollution stress; 1998; Edinburgh, Scotland. Water Air Soil Pollut. 116(1-2
47             special issue).
48       Sheppard, L. J.; Smith, R. I.; Cannell, M. G. R. (1989) Frost hardiness ofPicea rubens growing in spruce decline regions
49             of the Appalachians. Tree Physiol. 5: 25-37.
5 0       Showman, R. E. (1991) A comparison of ozone injury to vegetation during moist and drought years. J. Air Waste
51             Manage. Assoc. 41: 63-64.
52       Sild, E.; Pleijel, H.; Sellden, G. (1999) Response of subterranean clover (Trifolium subterraneum) to ozone in relation to
53             plant age and light conditions during exposure. New Phytol. 143: 315-321.
         January 2005                                     9-3 87       DRAFT-DO NOT QUOTE OR CITE

-------
  1       Simini, M; Skelly, J. M; Davis, D. D.; Savage, J. E.; Comrie, A. C. (1992) Sensitivity of four hardwood species to
  2             ambient ozone in north central Pennsylvania. Can. J. For. Res. 22: 1789-1799.
  3       Simon-Plas, F.; Rusterucci, C.; Milat, M.-L.; Humbert, C.; Montillet, J.-L.; Blein, J.-P. (1997) Active oxygen species
  4             production in tobacco cells elicited by cryptogein. Plant Cell Environ. 20: 1573-1579.
  5       Simpson, D.; Tuovinen, J.-P.; Emberson, L.; Ashmore, M.R. (2001) Characteristics of an ozone deposition module.
  6             Water Air Soil Pollut. FOCUS 1:  253-262.
  7       Simpson, D.; Tuovinen, J.-P.; Emberson, L.; Ashmore, M.R. (2003) Characteristics of an ozone deposition module II:
  8             Sensitivity analysis. Water Air Soil Pollut. 143: 123-137.
  9       Singh, S. N., ed. (2000) Trace gas emissions and plants. Dordrecht, The Netherlands: Kluwer Academic Publishers.
10       Sivasankar, S.; Sheldrick, B.; Rothstein, S. J. (2000) Expression of allene oxide synthase determines defense gene
11             activation in tomato. Plant Physiol. 122: 1335-1342.
12       Skarby, L.; Ro-Poulsen, H.; Wellburn, F. A. M.; Sheppard, L. J. (1998) Impacts of ozone onforests: aEuropean
13             perspective. NewPhytol.  139: 109-122.
14       Skeffington, R. A.; Sutherland, P. M. (1995) The effects of SO2 and O3 fumigation on acid deposition and foliar leaching
15             in the Liphook forest fumigation experiment. Plant Cell Environ. 18: 247-261.
16       Skelly, J. M. (1988) Forest decline versus tree decline the pathological considerations. Environ. Monit. Assess. 12:
17             23-27.
18       Skelly, J. M. (2000) Tropospheric ozone and its importance to forests and natural plant communities of the northeastern
19             United States. Northeast. Nat. 7: 221-236.
20       Skelly, J. M. (2003) Native plants as bioindicators of air pollutants: contributed papers to a symposium held in
21             conjunction with the 34th air pollution workshop. Environ. Pollut. 125: 1-2.
22       Smeulders, S. M.; Gorissen, A.; Joosten, N. N.; Van Veen, J. A. (1995) Effects of short-term ozone exposure on the
23             carbon economy of mature and juvenile Douglas firs [Pseudotsuga menziesii (Mirb.) Franco]. New Phytol. 129:
24             45-53.
25       Smirnoff, N.;  Conklin, P. L.; Loewus, F. A. (2001) Biosynthesis of ascorbate acid in plants: a renaissance. Annu. Rev.
26             Plant Physiol. Mol. Biol. 52: 437-467.
27       Smith, W. B. (2002) Forest inventory and analysis: a national inventory and monitoring program. Environ. Pollut. 116:
28             S233-S242.
29       Smith, G.; Greenhalgh, B.; Brennan, E.;  Justin, J. (1987) Soybean yield in New Jersey relative to ozone pollution and
30             antioxidant application. Plant Dis. 71: 121-125.
31       Smith, G.; Coulston, J.; Jepsen, E.; Prichard, T. (2003) A national ozone biomonitoring program - results from field
32             surveys of ozone sensitive plants in northeastern forests (1994-2000). Environ. Monit. Assess. 87: 271-291.
33       Soja, G.; Bid, M.; Gangl, H.; Redl, H. (1997) Ozone sensitivity of grapevine (Vitis vinifera L.): evidence for a memory
34             effect in a perennial crop plant? Phyton (Austria) 37: 265-270.
35       Soja, G.; Barnes, J. D.; Posch, M.; Vandermeiren, K.; Pleijel, H.; Mills, G. (2000) Phenological weighting of ozone
36             exposures in the calculation of critical levels for wheat, bean and plantain. Environ. Pollut. 109:  517-524.
37       Somers, G. L.; Chappelka, A. H.; Rosseau, P.; Renfro, J. R. (1998) Empirical evidence of growth decline related to
38             visible ozone injury. For. Ecol. Manage. 104: 129-137.
39       Spash, C. L. (1997) Assessing the economic benefits to agriculture from air pollution control. J. Econ. Surv. 11: 47-70.
40       Stan,H.-J.; Schicker, S. (1982) Effect of repetitive ozone treatment on bean plants - stress ethylene production and leaf
41             necrosis. Atmos. Environ. 16: 2267-2270.
42       Stan, H.-J.; Schicker, S.; Kassner, H. (1981) Stress ethylene evolution of bean plants - a parameter indicating ozone
43             pollution. Atmos. Environ. 15: 391-395.
44       Stark, R. W.; Cobb, F. W., Jr. (1969) Smog injury, root diseases and bark beetle damage in ponderosa pine. Calif. Agric.
45             23(9): 13-15.
46       Stark, R. W.; Miller, P. R.; Cobb, F. W., Jr.; Wood, D. L.; Parmeter, J. R., Jr. (1968) Photochemical oxidant injury and
47             bark beetle (Coleoptera: Scolytidae) infestation of ponderosa pine. I. Incidence of bark beetle infestation in injured
48             trees. Hilgardia 39: 121-126.
49       Staszak, J.; Grulke, N. E.; Prus-Glowacki, W. (2004) Genetic differences of Pinusponderosa (Dougl. Ex Laws) trees
50             tolerant and sensitive to ozone. Water Air Soil Pollut. 153: 3-14.
51       Stewart, C. A. (1998) Impact of ozone on the reproductive biology ofBrassica campestris L. and Plantago major L
52             [dissertation]. Leicestershire, United Kingdom: University of Loughborough, Department of Geology.
53       Stewart, C. A.; Black, V. J.; Black, C. R.; Roberts, J. A. (1996) Direct effects of ozone on the reproductive development
54             ofBrassica species. J. Plant Physiol. 148: 172-178.


         January 2005                                      9-388       DRAFT-DO NOT QUOTE OR CITE

-------
  1       Stitt, M. (1996) Plasmodesmata play an essential role in sucrose export from leaves: a step towards an integration of
  2             metabolic biochemistry and cell biology. Plant Cell 8: 565-571.
  3       Stribley, G. H.; Ashmore, M. R. (2002) Quantitative changes in twig growth pattern of young woodland beech (Fagus
  4             sylvatica L.) in relation to climate and ozone pollution over 10 years. For. Ecol. Manage. 157: 191-204.
  5       Sullivan, J. A.; Hale, B. A.; Ormrod, D. P. (1994) Impact of seasonal ozone exposure on yield and vegetative growth of
  6             primocane-fruiting raspberry. HortScience 29: 1059-1061.
  7       Suter, G. W., II. (1990) Endpoints for regional ecological risk assessments. Environ. Manage. (N. Y.) 14: 9-23.
  8       Swank, W. T.; Vose, J. M. (1990) Watershed-scale responses to ozone events in a Pinus strobus L. plantation. Water Air
  9             SoilPollut. 54: 119-133.
10       Taiz, L.; Zeiger, E. (2002) Plant physiology. 3rd ed. Sunderland, MA: Sinauer Associates, Publishers.
11       Takemoto, B. K.; Olszyk, D. M.; Johnson, A. G.; Parada, C. R. (1988) Yield responses of field-gown crops to acidic fog
12             and ambient ozone. J. Environ. Qual. 17: 192-197.
13       Takemoto, B. K.; Bytnerowicz, A.; Dawson, P. J.; Morrison, C. L.; Temple, P. J.  (1997) Effects of ozone on Pinus
14             ponderosa seedlings: comparison of responses in the first and second growing seasons of exposure. Can. J. For.
15             Res. 27: 23-30.
16       Takemoto, B. K.; Bytnerowicz, A.; Fenn, M. E. (2001) Current and future effects of ozone and atmospheric nitrogen
17             deposition on California's mixed conifer forests. For. Ecol. Manage. 144:  159-173.
18       Takeuchi,  Y.; Kubiske, M. E.; Isebrands, J. G.; Pregitzer, K. S.; Hendrey, G.; Karnosky, D. F. (2001) Photosynthesis,
19             light and nitrogen relationships in a young deciduous forest canopy under open-air CO2 enrichment. Plant Cell
20             Environ. 24: 1257-1268.
21       Tappel, A. L. (1972) Vitamin E and free radical peroxidation of lipids. In: Nair, P. P.; Kayden, H. J., eds. Vitamin E and
22             its role in cellular metabolism. New York, NY:  New York Academy of Sciences; pp. 12-28. (Annals of the New
23             York Academy of Sciences: v. 203).
24       Tausz, M.; Bytnerowicz, A.; Arbaugh, M. J.; Weidner, W.; Grill, D. (1999a) Antioxidants and protective pigments of
25             Pinus ponderosa needles at gradients of natural stresses and ozone in the San Bernardino Mountains in California.
26             Free Rad. Res. 31(suppl.): S113-S120.
27       Tausz, M.; Bytnerowicz, A.; Weidner, W.; Arbaugh, M. J.; Padgett, P.; Grill, D. (1999b) Pigments and phytoprotection
28             in needles of Pinus ponderosa trees with and without symptoms  of ozone injury. Phyton (Horn, Austria) 39:
29             219-224.
30       Tausz, M.; Bytnerowicz, A.; Weidner, W.; Arbaugh, M. J.; Padgett, P.; Grill, D. (1999c) Changes in free-radical
31             scavengers describe the susceptibility of Pinus ponderosa to ozone in southern Californian forests. Water Air Soil
32             Pollut. 116:249-254.
33       Tausz, M.; Bytnerowicz, A.; Arbaugh, M. J.; Wonisch, A.; Grill, D. (2001) Multivariate patterns of biochemical
34             responses of Pinus ponderosa trees at field plots in the San Bernardino Mountains, southern California. Tree
35             Physiol. 21:329-336.
36       Tausz, M.; Herbinger, K.; Posch, S.; Grulke, N. E. (2002) Antioxidant status of Pinus jeffreyi needles from mesic and
37             xeric microsites in early and late summer. Phyton (Horn, Austria) 42: 201-207.
3 8       Taylor, G. E., Jr.; Hanson, P. J. (1992) Forest trees and tropospheric ozone: role of canopy deposition and leaf uptake in
39             developing exposure-response relationships. Agric. Ecosyst. Environ. 42: 255-273.
40       Taylor, G. E., Jr.; Pitelka, L. F. (1992) Genetic diversity of plant populations and the role of air pollution. In: Barker, J.
41             R.;  Tingey, D.  T., eds. Air pollution effects on biodiversity. New York, NY: Van Nostrand Reinhold; pp. 111-130.
42       Taylor, G. E., Jr.; McLaughlin, S. B.; Shriner, D. S. (1982) Effective pollutant dose. In: Unsworth, M. H.; Ormrod, D. P.,
43             eds. Effects of gaseous air pollution in agriculture and horticulture. London, United Kingdom: Butterworth
44             Scientific; pp. 458-460.
45       Taylor, G. E., Jr.; Tingey, D.  T.; Ratsch, H. C. (1982) Ozone flux in Glycine max (L.) Merr: sites of regulation and
46             relationship to  leaf injury. Oecologia 53: 179-186.
47       Taylor, G. E., Jr.; Hanson, P. J.; Baldocchi, D. D. (1988a) Pollutant deposition to individual leaves and plant canopies:
48             sites of regulation and relationship to injury. In: Heck, W. W.; Taylor, O.  C.; Tingey, D. T., eds. Assessment of
49             crop loss from air pollutants. New York, NY: Elsevier Applied Science; pp. 227-257'.
50       Taylor, G. E., Jr.; Ross-Todd, B. M.; Gunderson, C. A. (1988b) Action of ozone on foliar gas exchange in Glycine max
51             L. Merr: a potential role for endogenous stress ethylene.  New Phytol. 110: 301-307.
52       Taylor, G. E.; Owens, J.  G.; Grizzard, T.; Selvidge, W. J. (1993) Atmosphere xcanopy interactions of nitric acid vapor in
53             loblolly-pine grown in open-top chambers. J. Environ. Qual. 22:  70-80.
         January 2005                                     9-3 89       DRAFT-DO NOT QUOTE OR CITE

-------
  1       Taylor, G. E., Jr.; Johnson, D. W.; Andersen, C. P. (1994) Air pollution and forest ecosystems: a regional to global
  2             perspective. Ecol. Appl. 4: 662-689.
  3       Taylor, G.; Ceulemans, R.; Ferris, R.; Gardner, S. D. L.; Shao, B. Y. (2001) Increased leaf area expansion of hybrid
  4             poplar in elevated CO2. From controlled environments to open-top chambers and to FACE. Environ. Pollut. 115:
  5             463-472.
  6       Temple, P. J. (1988) Injury and growth of Jeffrey pine and giant sequoia in response to ozone and acidic mist. Environ.
  7             Exp. Bot.  28: 323-333.
  8       Temple, P. J. (1999) Effects of ozone on understory vegetation in the mixed conifer forest. In: Miller, P. R.; McBride, J.
  9             R., eds. Oxidant air pollution impacts in the montane forests of southern California: a case study of the San
10             Bernardio Mountains. New York, NY: Springer; pp. 208-222. (Ecological studies: v. 134).
11       Temple, P. J.; Miller, P. R. (1994) Foliar ozone injury and radial growth of ponderosa pine. Can. J. For. Res. 24:
12             1877-1882.
13       Temple, P. J.; Riechers, G. H.; Miller, P. R. (1992) Foliar injury responses of ponderosa pine seedlings to ozone, wet and
14             dry acidic deposition, and drought. Environ. Exp. Bot. 32: 101-113.
15       Temple, P. J.; Riechers, G. H.; Miller, P. R.; Lennox, R. W. (1993) Growth responses of ponderosa pine to long-term
16             exposure to ozone, wet and dry acidic deposition, and drought. Can.  J. For. Res. 23: 59-66.
17       Teramura, A. H.; Sullivan J. H.; Lydon, J. (1990) Effects of UV-b radiation on soybean yield and seed quality: a 6-year
18             field study. Physiol. Plant. 80: 5-11.
19       Teskey, R. O.; Dougherty, P. M.; Wiselogel, A. E. (1991) Design and performance of branch chambers suitable for
20             long-term ozone fumigation of foliage in large trees. J. Environ. Qual. 20: 591-595.
21       Thalmair, M.; Bauw, G.; Thiel, S.; Dohring, T.; Langebartels, C.; Sandermann, H.,  Jr. (1996) Ozone and ultraviolet B
22             effects on the defense-related proteins p-l,3-glucanase and chitinase in tobacco. J. Plant Physiol.  148: 222-228.
23       Thorn, M.; Maretzki, A. (1985) Evidence for a plasmalemma redox system in sugarcane. Plant Physiol. 77: 873-876.
24       Tiedemann, A. V.; Firsching, K.  H. (1993) Effects of ozone exposure and leaf  age of wheat on infection processes of
25             Septoria nodomm Berk.. Plant Pathol. 42: 287-293.
26       Tiedemann, A. V.; Firsching, K.  H. (2000) Interactive effects of elevated ozone and carbon dioxide on growth and yield
27             of leaf rust-infected versus non-infected wheat. Environ. Pollut. 108: 357-363.
28       Tiedemann, A. V.; Weigel, H.-J.; Jager, H.-J. (1991) Effects of open-top chamber fumigations with ozone on three
29             fungal leaf diseases of wheat and the mycoflora of the phyllosphere. Environ. Pollut. 72: 205-224.
30       Tingey, D. T. (1984) The effects of ozone on plants in the United States. In: Grennfelt, P., ed. Proceedings of the
31             international workshop on the evaluation of the effects of photochemical oxidants on human health, agricultural
32             crops, forestry, materials and visibility; February-March; Goteborg, Sweden. Goteborg, Sweden: Swedish
33             Environmental Research Institute; pp. 60-75; report no. IVL-EM 1570.
34       Tingey, D. T.; Andersen, C. P. (1991) The physiological basis of differential plant sensitivity to changes in atmospheric
35             quality. In: Taylor, G. E.,  Jr.; Pitelka, L.  f.; Clegg, M. T., eds. Ecological genetics and air pollution. New York,
36             NY: Springer-Verlag.
37       Tingey, D. T.; Olszyk, D. M. (1985)  Intraspecific variability in metabolic responses to SO2. In: Winner, W. E.; Mooney,
38             H. A.; Goldstein, R. A., eds. Sulfur dioxide and vegetation: physiology, ecology, and policy issues. Stanford, CA:
39             Stanford University Press; pp. 178-205.
40       Tingey, D. T.; Taylor, G. E., Jr. (1982)  Variation in plant response to ozone: a  conceptual model of physiological events.
41             In: Unsworth, M. H.; Ormrod, D. P., eds. Effects of gaseous air pollution in agriculture and horticulture. London,
42             United Kingdom: Butterworth Scientific; pp. 113-138.
43       Tingey, D. T.; Fites, R. C.; Wickliff,  C. (1975)  Activity changes in selected enzymes from soybean leaves following
44             ozone exposure. Physiol. Plant. 33: 316-320.
45       Tingey, D. T.; Standley, C.; Field, R. W. (1976a) Stress ethylene evolution: a measure of ozone effects on plants. Atmos.
46             Environ.  10: 969-974.
47       Tingey, D. T.; Wilhour, R. G.; Standley, C. (1976b) The effect of chronic ozone exposures on the metabolite content of
48             ponderosa pine seedlings.  For. Sci. 22: 234-241.
49       Tingey, D. T.; Hogsett, W. E.; Lee, E. H. (1989) Analysis of crop loss for alternative ozone exposure indices. In:
50             Schneider, T.; Lee, S. D.;  Welters, G. J.  R.; Grant, L.  D., eds. Atmospheric ozone research and its policy
51             implications:  proceedings of the  3rd US-Dutch international symposium; May 1988; Nijmegen, The Netherlands.
52             Amsterdam, The Netherlands: Elsevier Science Publishers; pp. 219-227. (Studies in environmental science 35).
53       Tingey, D. T.; Hogsett, W. E.; Lee, E. H.; Herstrom, A. A.; Azevedo, S. H. (1991) An evaluation of various alternative
54             ambient ozone standards based on crop yield loss data. In: Berglund, R. L.; Lawson, D. R.; McKee, D. J., eds.


         January 2005                                     9-390       DRAFT-DO NOT QUOTE OR CITE

-------
  1             Tropospheric ozone and the environment: papers from an international conference; March 1990; Los Angeles,
  2             CA. Pittsburgh, PA: Air & Waste Management Association; pp. 272-288. (A&WMA transactions series no.
  3             TR-19).
  4       Tingey, D. T.; McVeety, B. D.; Waschmann, R.; Johnson, M. G.; Phillips, D. L.; Rygiewicz, P. T.; Olszyk, D. M. (1996)
  5             A versatile sun-lit controlled-environment facility for studying plant and soil processes. J. Environ. Qual. 25:
  6             614-625.
  7       Tingey, D. T.; Laurence, J. A.; Weber, J. A.; Greene, J.; Hogsett, W. E.; Brown, S.; Lee, E. H. (2001) Elevated CO2 and
  8             temperature alter the response of Pinusponderosa to ozone: A simulation analysis. Ecol. Appl.  11: 1412-1424.
  9       Tingey, D. T.; Rodecap, K. D.; Lee, E. H.; Hogsett, W. E.; Gregg, J. W. (2002) Pod development increases the ozone
10             sensitivity ofPhaseolus vulgaris. Water Air Soil Pollut.  139: 325-341.
11       Tingey, D. T.; Hogsett, W. E.; Lee, E. H.; Laurence, J. A. (2004) Stricter ozone ambient air quality standard has
12             beneficial effect on Ponderosa pine in California. Environ. Manage. 34: 397-405.
13       Tissue, D. T.; Griffin, K. L.; Ball, T. (1999) Photosynthetic adjustment in field-grown ponderosa pine trees after six years
14             of exposure to elevated CO2.  Tree Physiol. 19: 221-228.
15       Tjoelker, M. G.; Luxmoore, R. J. (1991) Soil nitrogen and chronic ozone stress influence physiology, growth and nutrient
16             status of Pinus taeda L. and Liriodendron tulipifera L. seedlings. New Phytol. 119: 69-81.
17       Tjoelker, M. G.; Volin, J. C.; Oleksyn, J.; Reich, P.  B. (1993) Light environment alters response to ozone stress in
18             seedlings of Acer saccharum Marsh, and hybrid Populus L. I. In situ net photosynthesis, dark respiration and
19             growth. New Phytol. 124: 627-636.
20       Tjoelker, M. G.; Volin, J. C.; Oleksyn, J.; Reich, P.  B. (1994) An open-air system for exposing forest-canopy branches to
21             ozone pollution. Plant Cell Environ. 17: 211-218.
22       Tjoelker, M. G.; Volin, J. C.; Oleksyn, J.; Reich, P.  B. (1995) Interaction of ozone pollution and light effects on
23             photosynthesis in a forest canopy experiment. Plant Cell Environ.  18: 895-905.
24       Tognetti, R.; Cherubini, P.; Innes, J. L. (2000) Comparative stem-growth rates of Mediterranean trees under background
25             and naturally enhanced ambient CO2 levels. New Phytol. 146: 59-74.
26       Tomlinson, H.; Rich, S. (1968) The  ozone resistance of leaves as related to their sulfhydryl and adenosine triphosphate
27             content. Phytopathology 58: 808-810.
28       Tomlinson, H.; Rich, S. (1971) Effect of ozone onsterols and sterol derivatives in bean leaves. Phytopathology 61:
29             1404-1405.
30       Tonneijck, A. E.  G. (1994) Effects of various ozone exposures on the susceptibility of bean leaves (Phaseolus vulgaris
31             L.) to Botrytis cinerea. Environ. Pollut. 85: 59-65.
32       Tonneijck, A. E.  G.; Leone, G. (1993) Changes in susceptibility of bean leaves (Phaseolus vulgaris) to Sclerotinia
33             sclerotiorum and Botrytis cinerea by pre-inoculative ozone exposures. Neth. J. Plant Pathol. 99: 313-322.
34       Tonneijck, A. E.  G.; VanDijk, C. J. (1994) Combined effects of ozone and ammonia on growth and yield of bean
35             (Phaseolus vulgaris L. cv. Pros). In: Fuhrer, J.; Achermann, B., eds. Critical levels for ozone: a UN-ECE
36             workshop report; November 1993; Bern, Switzerland. Liebefeld-Bern, Switzerland: Swiss Federal Research
37             Station for Agricultural Chemistry and Environmental Hygiene; pp. 282-286. (Schriftenreihe der FAC Liebefeld
38             no. 16).
39       Tonneijck, A. E.  G.; Van Dijk, C. J. (1997) Effects of ambient ozone on injury and yield of Phaseolus vulgaris at four
40             rural sites in The Netherlands as assessed by  using ethylenediurea (EDU). New Phytol.  135: 93-100.
41       Tonneijck, A. E.  G.; VanDijk, C. J. (1998) Responses of bean (Phaseolus vulgaris L. cv. Pros) to chronic ozone
42             exposure at two levels of atmospheric ammonia. Environ. Pollut. 99: 45-51.
43       Tonneijck, A. E.  G.; Van Dijk, C. J. (2002a) Assessing effects of ambient ozone on injury and yield  of bean with
44             ethylenediurea (EDU): three years of plant monitoring at four sites in The Netherlands. Environ. Monit. Assess.
45             77: 1-10.
46       Tonneijck, A. E.  G.; Van Dijk, C. J. (2002b) Injury  and growth response of subterranean clover to ambient ozone as
47             assessed by using ethylenediurea (EDU): three years of plant monitoring at four sites in The Netherlands. Environ.
48             Exper. Bot. 48:  33-41.
49       Topa, M. A.; Vanderklein, D. W.; Corbin, A. (2001) Effects of elevated ozone and low light on diurnal and seasonal
50             carbon gain in sugar maple. Plant Cell Environ. 24: 663-677'.
51       Treshow, M.; Stewart, D. (1973) Ozone sensitivity of plants in natural communities. Biol. Conserv. 5:  209-214.
52       Tuovinen, J.-P. (2002) Assessing vegetation exposure to ozone: is it possible to estimate AOT40 by passive sampling.
53             Environ. Pollut. 119: 203-214.
         January 2005                                     9-391       DRAFT-DO NOT QUOTE OR CITE

-------
  1       U.S. Congress. (2002) Food security and rural investment act of 2002, PL 107-171, May 13, 2002. Washington, DC:
  2             U.S. Government Printing Off ice.
  3       U.S. Environmental Protection Agency. (1986) Air quality criteria for ozone and other photochemical oxidants. Research
  4             Triangle Park, NC: Office of Health and Environmental Assessment, Environmental Criteria and Assessment
  5             Office; report nos. EPA-600/8-84-020aF-eF. 5v. Available from: NTIS, Springfield, VA; PB87-142949.
  6       U.S. Environmental Protection Agency. (1992) Summary of selected new information on effects of ozone on health and
  7             vegetation: supplement to 1986 air quality criteria for ozone and other photochemical oxidants. Research Triangle
  8             Park, NC: Office of Health and Environmental Assessment, Environmental Criteria and Assessment Office; report
  9             no. EPA/600/8-88/105F. Available from: NTIS, Springfield, VA; PB92-235670.
10       U.S. Environmental Protection Agency. (1993) Air quality criteria for oxides of nitrogen. Research Triangle Park, NC:
11             Office of Health and Environmental Assessment, Environmental Criteria and Assessment Office; report nos.
12             EPA/600/8-9!/049aF-cF. 3v. Available from: NTIS, Springfield, VA; PB95-124533, PB95-124525, and
13             PB95-124517.
14       U.S. Environmental Protection Agency. (1996a)  Air quality criteria for ozone and related photochemical oxidants.
15             Research Triangle Park, NC: Office of Research and Development; report nos. EP A/600/AP-93/004aF-cF. 3v.
16             Available from: NTIS, Springfield, VA; PB96-185582, PB96-185590, and PB96-185608. Available:
17             http://cfpub2.epa.gov/ncea/.
18       U.S. Environmental Protection Agency. (1996b)  Review of national ambient air quality standards  for ozone: assessment
19             of scientific and technical information. OAQPS staff paper. Research Triangle Park, NC: Office of Air Quality
20             Planning and Standards; report no. EPA/452/R-96/007. Available from: NTIS, Springfield, VA; PB96-203435.
21             Available: http://www.epa.gov/ttn/naaqs/standards/ozone/s_o3_pr_sp.html (9 September 2003).
22       U.S. Environmental Protection Agency. (2003) Technology Transfer Network: Air Quality System (AQS).  Washington,
23             DC: Office of Air and Radiation. Available: http://www.epa.gov/ttn/airs/airsaqs/ [1 April, 2004].
24       United Nations Economic Commission for Europe, UNECE. (1988) ECE critical levels workshop; March; Bad
25             Harzburg, Germany [final report]. Geneva, Switzerland: United Nations Economic Commission for Europe.
26       United Nations Environment Programme (UNEP). (1993) Environmental data report 1993-94. Oxford, United Kingdom:
27             Blackwell Publishers.
28       United Nations Environment Programme (UNEP). (1999) Geo-2000: global environmental outlook 2000. London,
29             United Kingdom: Earthscan Publishers. Available: http://www.unep.org/geo2000/english/index.htm (26 July
30             2002).
31       Unsworth, M. H.; Hogsett, W. E. (1996) Combined effects of changing CO2, temperature, UV-B radiation and O3 on
32             crop growth. In: Bazzaz, F.; Sombroek, W., eds. Global climate change and agricultural production.  Chichester,
33             United Kingdom: John Wiley & Sons; pp. 171-197. Available:
34             http://www.fao.org/docrep/W5183E/w5183eOO.htm [28 August, 2003].
35       Unsworth, M. H.; Heagle, A. S.; Heck, W. W. (1984a) Gas exchange in open-top field chambers -1. measurement and
36             analysis of atmospheric resistances to gas exchange. Atmos. Environ. 18: 373-380.
37       Unsworth, M. H.; Heagle, A. S.; Heck, W. W. (1984b) Gas exchange in open-top field chambers - II. resistances to
38             ozone uptake by soybeans. Atmos. Environ. 18: 381-385.
39       Utriainen, J.; Holopainen, T. (2001) Nitrogen availability modifies the ozone responses of Scots pine seedlings exposed
40             in an open-field system. Tree Physiol. 21: 1205-1213.
41       Utriainen, J.; Janhunen, S.; Helmisaari, H. S.; Holopainen, T. (2000) Biomass allocation, needle structural characteristics
42             and nutrient composition in Scots pine seedlings exposed to elevated CO2 and O3 concentrations. Trees (Berlin,
43             Ger.) 14: 475-484.
44       Vacek, S.; Bastl, M.; Leps, J. (1999) Vegetation  changes in forests of the Krkonose Mountains over a period of air
45             pollution stress (1980-1995). Plant Ecol. 143: 1-11.
46       Van Camp, W.; Willekens, H.; Bowler, C.; Van Montagu, M.; Inze, D.; Reupold-Popp, P.; Sandermann, H., Jr.;
47             Langebartels, C. (1994) Elevated levels of superoxide dismutase protect transgenic plants against ozone damage.
48             Biotechnology 12: 165-168.
49       Van De Staaij, J. W. M.; Lenssen, G. M.; Stroetenga, M.; Rozema, J. (1993) The combined effects of elevated CO2 levels
50             and UV-B radiation on growth characteristics ofElymus athericus (= E. pycnanthus). Vegetatio 104/105: 433-439.
51       Van De Staaij, J. W. M.; Tonneijck, A. E. G.; Rozema, J. (1997) The effect of reciprocal treatments with ozone and
52             ultraviolet-B radiation on photosynthesis and growth of perennial grass Elymus athericus. Environ. Pollut. 97:
53             281-286.
         January 2005                                    9-392      DRAFT-DO NOT QUOTE OR CITE

-------
  1       Van Den Driessche, R.; Langebartels, C. (1994) Foliar symptoms, ethylene biosynthesis and water use of young Norway
  2             spruce (Picea abies (L.) Karst.) exposed to drought and ozone. Water Air Soil Pollut. 78: 153-168.
  3       Van Oijen, M; Ewart, F. (1999) The effects of climatic variation in Europe on the yield response of spring wheat cv.
  4             Minaret to elevated CO2 and O3: an analysis of open-top chamber experiments by means of two crop growth
  5             simulation models. Eur. J. Agron. 10: 249-264.
  6       Van Oijen, M.; Schapendonk, A. H. C. M.; Jansen, M. J. H.; Pot, C. S.; Maciorowski, R. (1999) Do open-top chambers
  7             overestimate the effects of rising CO2 on plants? An analysis using spring wheat. Glob. Change Biol. 5: 411-421.
  8       Vanacker, H.; Carver, T. L. W.; Foyer, C. H. (2000) Early H2O2 accumulation in mesophyll cells leads to induction of
  9             glutathione during the hyper-sensitive response in the barley-powdery mildew interaction. Plant Physiol. 123:
10             1289-1300.
11       VanderHeyden, D.; Skelly, J.; Innes, J.; Hug, C.; Zhang, J.; Landolt, W.; Bleuler, P. (2001) Ozone exposure thresholds
12             and foliar injury on forest plants in Switzerland. Environ. Pollut. Ill: 321-331.
13       Vandermeiren, K.; De Temmerman, L.; Hookham, N. (1995) Ozone sensitivity of Phaseolus vulgaris in relation to
14             cultivar differences, growth stage and growing conditions. Water Air Soil Pollut. 85: 1455-1460.
15       Vandermeiren, K.; Black, C.; Lawson, T.; Casanova, M. A.; Ojanpera, K. (2002) Photosynthetic and stomatal responses
16             of potatoes grown under elevated CO2 and/or O3—results from the European CHIP-programme. Eur. J. Agron. 17:
17             337-352.
18       Vann, D. R.; Strimbeck, G. R.; Johnson, A. H. (1995) Effects of mist acidity and ambient ozone removal on montane red
19             spruce. Tree Physiol. 15: 639-647.
20       Volin, J. C.; Reich, P. B.; Givnish, T. J. (1998) Elevated carbon dioxide ameliorates the effects of ozone on
21             photosynthesis and growth: species respond similarly regardless of photosynthetic pathway or plant functional
22             group. New Phytol. 138: 315-325.
23       Volk, M.; Geissmann, M.; Blatter, A.; Contat, F.; Fuhrer, J. (2003) Design and performance of a free-air exposure system
24             to study long-term effects of ozone on grasslands. Atmos. Environ. 37: 1341-1350.
25       Vollenweider, P.; Woodcock, H.; Kelty, M. J.; Hofer, R.-M. (2003a) Reduction of stem growth and site dependency of
26             leaf injury in Massachusetts black cherries exhibiting ozone symptoms. Environ. Pollut. 125: 467-480.
27       Vollenweider, P.; Ottiger, M.; Gunthardt-Goerg,  M. S. (2003b) Validation of leaf ozone symptoms in natural vegetation
28             using  microscopical methods. Environ. Pollut. 124: 101-118.
29       Von Gunten, U. (2003) Ozonation of drinking water: part I. Oxidation kinetics and product formation. Water Res. 37:
30             1443-1467.
31       Vorne, V; Ojanpera, K.; De Temmerman, L.; Bindi, M.; Hogy, P.; Jones, M. B.; Lawson, T.; Persson, K. (2002) Effects
32             of elevated carbon dioxide and ozone on potato tuber quality in the European multiple-site experiment
33             'CHIP-projecf. Eur. J. Agron. 17: 369-381.
34       Vozzo, S. F.; Miller, J. E.; Pursley, W. A.; Heagle, A. S. (1995) Effects of ozone and water deficit on field-grown
35             soybean: I. Leaf gas exchange. J. Environ. Qual. 24: 663-670.
36       Waite, C. E.; DeHayes, D. H.; Rebbeck, J.; Schier, G. A.; Johnson, A.  H. (1994) The influence of elevated ozone  on
37             freezing tolerance of red spruce seedlings. New Phytol. 126: 327-335.
38       Walcek, C. J.; Yuan, H.-H.; Stockwell, W. R. (1997) The influence of aqueous-phase chemical reactions on ozone
39             formation in polluted and nonpolluted clouds. Atmos. Environ. 31: 1221-1237.
40       Wallin, G.; Skarby, L.; Sellden, G. (1992) Long-term exposure of Norway spruce, Picea abies (L.) Karst., to ozone in
41             open-top chambers. 3. Effects on the light response of net photosynthesis in shoots of different ages. New Phytol.
42             121:387-394.
43       Wallin, G.; Karlsson, P. E.; Sellden, G.; Ottosson, S.; Medin, E.-L.; Pleijel, H.; Skarby, L.  (2002) Impact of fouryears
44             exposure to different levels of ozone, phosphorus and drought on chlorophyll, mineral nutrients, and stem volume
45             of Norway spruce, Picea abies. Physiol. Plant. 114: 192-206.
46       Wang, D.; Hinckley, T. M.; Gumming, A. B.; Braatne, J. (1995) A comparison of measured and modeled ozone uptake
47             into plant leaves. Environ. Pollut. 89: 247-254.
48       Ward, E. R.; Uknes, S. J.; Williams, S. C.; Dincher, S. S.; Wiederhold, D. L.; Alexander, D. C.; Ahi-Goy, P.; Metraux,
49             J.-P.; Ryals, J. A. (1991) Coordinate gene  activity in response to agents that induce systemic acquired resistance.
50             Plant Cell 3: 1084-1094.
51       Warwick, K. R.; Taylor, G. (1995) Contrasting effects of tropospheric  ozone on five native herbs which coexist in
52             calcareous grassland. Global Change Biol. 1: 143-151.
53       Weber, J.; Grulke, N. E. (1995) Effects of air pollution on stem growth. In: Gartner, B., ed. Stems and trunks in plant
54             form and function. San Diego,  CA: Academic Press; pp. 343-364.


         January 2005                                     9-393        DRAFT-DO NOT QUOTE OR CITE

-------
  1       Weber, J. A.; Clark, C. S.; Hogsett, W. E. (1993) Analysis of the relationships among O3 uptake, conductance, and
  2             photosynthesis in needles of Pinusponderosa. Tree Physiol. 13: 157-172.
  3       Weiner, J.; Fredro-Boniecki, S.; Reed, D.; Maclean, A.; Strong, M. (1997) NiepoLomice forest-A GIS analysis of
  4             ecosystem response to industrial pollution. Environ. Pollut. 3:  381-388.
  5       Weinstein, D. A.; Yanai, R. D. (1994) Integrating the effects of simultaneous multiple stresses on plants using the
  6             simulation model TREGRO. J. Environ. Qual. 23: 418-428.
  7       Weinstein, D. A.; Samuelson, L. J.; Arthur, M. A. (1998) Comparison of the response of red oak (Quercus rubra)
  8             seedlings and mature trees to ozone exposure using simulation modeling. Environ. Pollut. 102: 307-320.
  9       Weinstein, D. A.; Gollands, B.; Retzlaff, W. A. (2001) The effects of ozone on a lower slope forest of the Great Smoky
10             Mountain National Park: simulations linking an individual tree model to a stand model. For.  Sci. 47: 29-42.
11       Weinstein, D. A.; Woodbury, P. B.; Gollands, B.; King, P.; Lepak, L.; Pendleton, D. (2002) Assessment of effects of
12             ozone on forest resources in the southern Appalachian mountains. Asheville, NC: Southern Appalachian
13             Mountains Initiative. Available: http://cycas.cornell.edu/ebp/pmg.html [19 March, 2003].
14       Welfare, K.; Flowers, T. J.; Taylor, G.; Yeo, A. R.  (1996) Additive and antagonistic effects of ozone and salinity on the
15             growth, ion contents and gas exchange of five varieties of rice (Oryza sativa L.). Environ. Pollut. 92: 257-266.
16       Wellburn, A. R. (1990)  Why are atmospheric oxides of nitrogen usually phytotoxic and not alternative fertilizers? New
17             Phytol. 115:395-429.
18       Wellburn, A. (1994) Air pollution and climate change: the biological impact. Harlow, Essex, United Kingdom: Longman
19             Scientific & Technical.
20       Wellburn, F. A. M.; Wellburn, A. R. (1994) Atmospheric ozone affects carbohydrate allocation and winter hardiness of
21             Pinus halepensis (Mill.) J. Exp. Bot. 274: 607-614.
22       Werner, H.; Fabian, P. (2002) Free-air fumigation of mature trees. A  novel system for controlled ozone enrichment in
23             grown-up beech and spruce canopies. Environ.  Sci. Pollut. Res. 9: 117-121.
24       Wesely, M. L.  (1989) Parameterization of surface resistances to gaseous dry deposition in regional-scale numerical
25             models. Atmos. Environ. 23: 1293-1304.
26       Westenbarger, A. D.; Frisvold, G. B. (1994) Agricultural exposure to ozone and acid precipitation. Atmos. Environ. 28:
27             2895-2907.
28       Westman, W. E. (1979) Oxidant effects on Californian coastal sage scrub.  Science (Washington, DC) 205:  1001-1003.
29       Westman, W. E. (1981) Factors influencing the distribution of species of Californian coastal sage scrub. Ecology 62:
30             439-455.
31       Weubbles, D. J.; Jain, A.; Edmonds, J.; Harvey, D.; Hayhoe, K. (1999) Global change: state of the science. Environ.
32             Pollut. 100: 57-86.
33       Wheeler, T. R.; Hong, T. D.; Ellis, R. H.; Batts, G. R.; Morrison, J. I. L.; Hadley, P. (1996) The duration and rate of grain
34             growth, and harvest index, of wheat (Triticum aestivum L.) in  response to temperature and CO2. J. Exper. Bot. 47:
35             623-630.
36       Whitfield, C. P.; Davison, A. W.; Ashenden, T. W. (1996a) Interactive effects of ozone and soil volume onPlantago
37             major. New Phytol. 134: 287-294.
38       Whitfield, R. G.; Biller, W. F.; Jusko, M. J.; Keisler, J. M. (1996b) A probabilistic assessment of health risks associated
39             with short-term exposure to tropospheric ozone. Argonne, IL:  Argonne National Laboratory, Decision and
40             Information Sciences Division; report no. ANL/DIS-3.
41       Whitfield, C. P.; Davison, A. W.; Ashenden, T. W. (1997) Artificial selection and heritability of ozone resistance in two
42             populations ofPlantago major. New Phytol. 137: 645-655.
43       Whitfield, C. P.; Davison, A. W.; Ashenden, T. W. (1998) The effects of nutrient limitation on the response ofPlantago
44             major to ozone. New Phytol. 140: 219-230.
45       Whittaker, J. B. (1994) Interactions between insects and air pollutants. In: Alscher, R. G.; Wellburn, A. R., eds. Plant
46             responses to the gaseous environment. Molecular, metabolic and physiological aspects. London, United Kingdom:
47             Chapman and Hall; pp. 365-384.
48       Wickman, B. E. (1992) Forest Health in the Blue Mountains: the influence of insects and diseases. Portland OR: U.S.
49             Department of Agriculture, Forest Service, Pacific Northwest Research Station; general technical report
50             PNW-GTR-295.  (Quigley, T. M., ed. Forest health in the Blue Mountains: science perspectives).
51       Wieser, G. (1997) Ozone impact on photosynthetic capacity of mature and young Norway spruce (Picea abies (L.)
52             Karst): External versus internal exposure. Phyton (Austria) 37: 297-302.
53       Wieser, G.; Havranek, W. M. (1993) Ozone uptake in the sun and shade crown of spruce: quantifying the physiological
54             effects of ozone exposure. Trees (Berlin, Ger.) 7: 227-232.


         January 2005                                     9-394       DRAFT-DO NOT QUOTE OR CITE

-------
  1       Wieser, G.; Havranek, W. M. (1995) Environmental control of ozone uptake inLarix decidua Mill.: a comparison
  2             between different altitudes. Tree Physiol. 15: 253-258.
  3       Wieser, G.; Hasler, R.; Gotz, B.; Koch, W.; Havranek, W. M. (1999) Seasonal ozone uptake of mature evergreen
  4             conifers at different altitudes. Phyton (Austria) 39: 233-240.
  5       Wieser, G.; Tegischer, K.; Tausz, M.; Haberle, K.-H.; Grams, T. E. E.; Matyssek, R. (2002) Age effects on Norway
  6             spruce (Picea abies) susceptibility to ozone uptake: a novel approach relating stress avoidance to defense. Tree
  7             Physiol. 22: 583-590.
  8       Wilbourn, S.; Davison, A. W.; Ollerenshaw, J. H. (1995) The use of an unenclosed field fumigation system to determine
  9             the effects of elevated ozone on a grass-clover mixture. New Phytol. 129: 23-32.
10       Willekens, H.; Van Camp, M.; Van Montagu, M.; Inze, D.; Langebartels, C.; Sandermann, H., Jr. (1994) Ozone, sulfur
11             dioxide and ultraviolet B have similar effects on mRNA accumulation o fantioxidant genes in Nicotiana
12             plumbaginifolia L. Plant Physiol. 106:1007-1014.
13       Williams, T. P. W. (1994) Use of passive monitors to assess plant bioindicators for ground-level ozone. New Brunswick,
14             Canada: University of New Brunswick.
15       Williams, J.; Bulman, M. P.; Neill, S. J. (1994) Wilt-induced aba biosynthesis, gene expression and down regulation of
16             rbcS mRNA levels mArabidopsis thaliana. Physiol. Plant. 91:  177-182.
17       Wiltshire, J. J. J.; Wright, C. J.; Unsworth, M. H. (1992) A new method for exposing mature trees to ozone episodes. For.
18             Ecol. Manage. 51: 115-120.
19       Wiltshire, J. J. J.; Unsworth, M. H.; Wright, C. J. (1994) Seasonal changes in water use of ash trees exposed to ozone
20             episodes. New Phytol. 127: 349-354.
21       Wiltshire, J. J. J.; Wright, C. J.; Colls, J. J.; Craigon, J.; Unsworth, M. H. (1996) Some foliar characteristics of ash trees
22             (Fraxinus exelsior) exposed to ozone episodes. New Phytol. 134:  623-630.
23       Winner, W. E.; Lefohn, A. S.; Cotter, I. S.; Greitner, C. S.; Nellessen, J.; McEvoy, L. R., Jr.; Olson, R. L.; Atkinson, C.
24             J.; Moore, L. D. (1989) Plant responses to elevational gradients of O3 exposures in Virginia. Proc. Natl. Acad. Sci.
25             U. S. A. 86: 8828-8832.
26       Winner, W. E.; Coleman, J. S.; Gillespie, C.; Mooney, H. A.; Pell, E. J. (1991) Consequences of evolving reistance to air
27             pollutants. In: Taylor, G. E.; Pitelka, L. F.; Clegg, M. T., eds. Ecological genetics and air pollution. New York,
28             NY: Springer-Verlag; pp. 177-202.
29       Wolf, J.; Van Oijen, M. (2003) Model simulation of effects of changes in climate and atmospheric CO2 and O3 on tuber
30             yield potential of potato (cv. Bintje) in the European Union. Agric. Ecosyst. Environ. 94: 141-157.
31       Wolff, K.; Morgan-Richards, M.; Davison, A. W. (2000) Patterns of molecular genetic variation inPlantago major and
32             P. intermedia in relation to ozone resistance. New Phytol. 145:  501-509.
33       Wolters, J. H. B.; Martens, M. J. M.  (1987) Effects of air pollutants on pollen. Bot. Rev. 53: 372-414.
34       Woodbury, P. B.; Laurence, J. A.; Hudler,  G. W. (1994) Chronic ozone exposure increases the susceptibility of hybrid
3 5             Populus to disease caused by  Septoria musiva. Environ. Pollut. 86: 109-114.
36       Wu, Y.-X.; Tiedemann, A. V. (2002) Impact of fungicides on active oxygen species and antioxidant enzymes in spring
37             barley (Hordeum vulgare L.)  exposed to ozone. Environ. Pollut. 116: 37-47.
38       Wulff, A.; Hanninen, O.; Tuomainen, A.; Karenlampi, L. (1992) A method for open-air exposure of plants to ozone.
39             Ann. Bot. Fennici. 29: 253-262.
40       Wullschleger, S. D.; Hanson, P. J.; Edwards, G. S. (1996) Growth and maintenance respiration in leaves of northern red
41             oak seedlings and mature trees after 3 years of ozone exposure. Plant Cell Environ. 19: 577-584.
42       Wustman, B. A.; Oksanen, E.; Karnosky, D. F.; Noormets, A.; Isebrands, J. G.; Pregitzer, K. S.; Hendrey, G. R.; Sober,
43             J.; Podila, G. K. (2001) Effects of elevated CO2 and O3 on aspen clones varying in O3 sensitivity: can CO2
44             ameliorate the harmful effects of O3? Environ. Pollut. 115: 473-481.
45       Yalpani, N.; Enyedi, A. J.; Leon, J.;  Raskin, I. (1994) Ultraviolet light and ozone stimulate accumulation of salicyclic
46             acid, pathogenesis-related proteins and virus resistance in tobacco. Planta 193: 372-376.
47       Yonekura, T.; Dokiya, Y.; Fukami, M.; Izuta, T. (2001) Effects of ozone and/or soil water stress on growth and
48             photosynthesis ofFagaus crenata seedlings. Water Air Soil Pollut. 130: 965-970.
49       Young, T. F.; Sanzone, S., eds. (2002) A framework for assessing and reporting on ecological condition: an SAB report.
50             Washington, DC: U.S. Environmental Protection Agency, Science Advisory Board; report no.
51             EPA-SAB-EPEC-02-009. Available: http://www.epa.gov/sab/pdf/epec02009.pdf [9 December, 2003].
52       Younglove, T.; McCool, P. M.; Musselman, R. C.; Kahl, M. E. (1994) Growth-stage dependent crop yield response to
53             ozone exposure. Environ. Pollut. 86: 287-295.
         January 2005                                     9-395       DRAFT-DO NOT QUOTE OR CITE

-------
  1       Yun, S.-C.; Laurence, J. A. (1999a) The response of clones ofPopulus tremuloides differing in sensitivity to ozone in the
  2             field. New Phytol. 141: 411-421.
  3       Yun, S.-C.; Laurence, J. A. (1999b) The response of sensitive and tolerant clones ofPopulus tremuloides to dynamic
  4             ozone exposure under controlled environmental conditions. New Phytol. 143: 305-313.
  5       Yunus, M; Iqbal, M. (1996) Plant response to air pollution. Chichester, United Kingdom: John Wiley & Sons Ltd.
  6       Yuska, D. E.; Skelly, J. M.; Ferdinand, J. A.; Stevenson, R. E.; Savage, J. E.; Mulki, J.; Mines, A. (2003) Use of
  7             bioindicators and passive sampling devices to evaluate ambient ozone concentrations in north central
  8             Pennsylvania. Environ. Pollut. 125:  71-80.
  9       Zeller, K.; Nikolov, N. (2000) Quantifying simultaneous fluxes of ozone, carbon dioxide and water vapor above a
10             subalpine forest ecosystem. Environ. Pollut. 107: 1-20.
11       Zhang, X.; Zhang, L.; Dong, F.; Gao, J.; Galbraith, D. W.; Song, C.-P. (2001) Hydrogen peroxide is involved in abscisic
12             acid-induced stomatal closure in Viciafaba. Plant Physiol. 126: 1438-1448.
13       Zheng, Y.; Lyons, T.; Ollerenshaw, J. H.; Barnes, J. D. (2000) Ascorbate in the leaf apoplast is a factor mediating ozone
14             resistance in Plantago major. Plant Physiol. Biochem. (Paris, Fr.) 3 8: 403 -411.
15       Zinser, C.; Jungblut, T.: Heller, W.; Seidlitz, H. K.; Schnitzler, J.-P.; Ernst, D.; Sandermann, H., Jr. (2000) The effect of
16             ozone in Scots pine (Pinus sylvestris L.): gene expression, biochemical changes and interactions with UV-B
17             radiation. Plant Cell Environ. 23: 975-982.
18       Ziska, L.; Bunce, J. A. (1997)  The role of temperature in determining the stimulation of CO2 assimilation at elevated
19             carbon dioxide concentrations in soybean seedlings. Physiol. Plant. 100:  126-132.
20
         January 2005                                     9-396       DRAFT-DO NOT QUOTE OR CITE

-------
 i      10.  TROPOSPHERIC OZONE EFFECTS ON UV-B FLUX
 2                  AND CLIMATE CHANGE PROCESSES
 3
 4
 5      10.1  INTRODUCTION
 6          In addition to exerting direct effects on human health and vegetation/ecosystems,
 7     tropospheric ozone (O3) is involved in determining ground-level flux of solar ultraviolet (UV)
 8     radiation and, also, in other processes that alter the earth's radiative balance and contribute to
 9     resulting climate change.  This chapter first discusses the role of tropospheric ozone in
10     determining surface-level UV flux and, then, secondly, discusses tropospheric O3 involvement in
11     global climate change.
12
13
14      10.2   THE ROLE OF TROPOSPHERIC OZONE IN DETERMINING
15            GROUND-LEVEL UV-B FLUX
16          Stratospheric O3 plays a crucial role in reducing the exposure of living organisms to solar
17     ultraviolet radiation.  The stratosphere, the atmospheric layer immediately above the
18     troposphere, is roughly 50 km thick and possesses 90% or more of the atmosphere's O3.
19     Ozone depletion due to the release of long-lived anthropogenic chlorinated and fluorinated
20     hydrocarbons was discovered over the course of the 1970s and early 1980s and led to
21     establishment of an international treaty for the protection of stratospheric O3, the 1987 Montreal
22     Protocol on Substances that Deplete the Ozone Layer.
23          Active research has been underway into the effects of increased UV radiation on
24     ecosystems and on human health since the discovery of the seasonal polar "Ozone Hole" and the
25     declining stratospheric O3 concentrations in the midlatitudes. An outcome of this effort is a body
26     of literature that also describes the effects of tropospheric pollutants, PM, and O3, on ground-
27     level UV radiation flux.  The Montreal Protocol requires routine review of the latest scientific
28     information available on the status of the O3 layer and of UV radiation levels at the earth's
29     surface. The World Meteorology Organization (WMO) and UN Environmental Program
30     (UNEP) are responsible for assessing the state of the science regarding the O3 layer and for
       January 2005                             10-1        DRAFT-DO NOT QUOTE OR CITE

-------
 1      reporting on trends in surface UV radiation levels. The latest WMO/UNEP assessment was
 2      published in 2003 (WMO/UNEP, 2002).
 3           This section describes the current level of scientific understanding of the factors
 4      influencing the flux of UV radiation at the earth's surface, including geophysical factors,
 5      tropospheric O3, PM, and cloud cover, with reference to the WMO/UNEP assessment and the
 6      peer-reviewed literature.  Factors influencing the degree of human exposure to UV-B and the
 7      resulting health effects are also discussed.
 8
 9      10.2.1  Factors  Governing Ultraviolet Radiation Flux at the Earth's Surface
10      10.2.1.1  UV Radiation: Wavelengths, Energies and Depth of Atmospheric Penetration
11           Designations for portions of the electromagnetic spectrum have evolved over time and are
12      usually associated with general function or effects caused by photons in a given wavelength
13      range. Gamma rays, at or below 0.1 nm in wavelength, are especially damaging high energy
14      photons emitted during radioactive decay and by stellar activity.  Radiowaves, wavelengths
15      greater than 108 nm, are very low in energy and function as carriers for broadcast
16      communications.  The energy possessed by any photon is inversely proportional to its
17      wavelength.
18           The wavelengths ranging between 50 and 400 nm in length are denoted "ultraviolet."
19      Solar radiation of wavelengths < 280 nm, including UV-C (200 to 280 nm), is almost entirely
20      blocked by Earth's upper atmosphere due to photoionization and photo-dissociation  processes.
21      Figure 10-1 shows a comparison between solar flux above the atmosphere and the flux at the
22      earth's surface. The flux of solar radiation between 280 and 320 nm (UV-B) is absorbed or
23      scattered in part within the atmosphere, while radiation between 320 and 400 nm (UV-A) can be
24      scattered but are not absorbed by gases to any meaningful degree. UV-B and UV-A photons
25      contain the necessary levels of energy needed to break (photolyze) chemical bonds.  Both UV-A
26      and UV-B are associated with human health- and ecosystem-damaging effects. Because UV-B
27      is more energetic, it is potentially capable of producing more biological damage than UV-A.
28      UV-A comprises 90% of the UV radiation that  reaches the earth's surface, yet is 1000 times less
29      damaging to keratinocytes (skin cells), than UV-B (Hildesheim, 2004).
30
31

        January 2005                             10-2       DRAFT-DO NOT QUOTE OR CITE

-------
                   0)
                   o
                      1.0000
                   I  0.1000
                   es
                   E
                   x  0.0100
                      0.0010
                      0.0001
                                                Solar Flux
                          290
                                        300           310           320
                                               Wavelength (nm)
                                                                                 330
       Figure 10-1.  Extraterrestrial solar flux measured by the satellite UARS SOLSTICE
                     instrument (dotted line). The dashed line represents calculated atmospheric
                     transmittance and the solid line is the calculated absolute flux of UV
                     radiation for a solar zenith angle of SOdeg, total column O3 of 275 DU, and
                     a surface reflectivity of 8%. The fine structure on the surface flux trace
                     results from Fraunhofer lines (absorption specific wavelengths within the
                     solar atmosphere).
       Source: Krotkovetal. (1998).
 1     10.2.1.2  Temporal Variations in Solar Flux
 2          The magnitude of the solar radiation flux entering the atmosphere is dependent upon long-
 3     term solar activity, sunspot cycle (11 years), solar rotation (27 days) and the position of the earth
 4     in its orbit around the sun. A variety of changes in solar irradiance can be found in the historical
 5     data, beginning in 1700 and leading up to the present. Fligge and Solanki (2000) concluded that
 6     solar irradiation changes on time scales of days to centuries can be attributed to variations in
 7     solar magnetic features.  Since the last Maunder minimum in 1700, solar irradiance has
 8     increased very slightly, at approximately 3.0% for wavelengths < 300 nm and at -1.3% for the
 9     UV-B and UV-A range.  Including visible wavelengths, Fligge and Solanki (2000) estimated that
10     the overall increase in solar irradiance was -0.3%. Rozema et al. (2001) pointed out that any
       January 2005
10-2
DRAFT-DO NOT QUOTE OR CITE

-------
 1      increase in wavelengths < 300 nm (UV-C) would initiate additional O3 formation in the
 2      stratosphere, therefore increasing its UV radiation absorptive capacity.
 3           Solar rotation and sunspot activity have the greatest effect on the radiation flux that
 4      originates in the highest levels of the solar atmosphere.  The amplitude of the associated cyclical
 5      changes in solar shortwave radiation flux follows an inverse relationship between the photon's
 6      wavelength and the solar altitude at which it was emitted. The maximum level of radiation
 7      (solar-max) differs from the minimum (solar-min) by as much as 10% for wavelengths near
 8      160 nm.  This peak-to-trough difference declines to around 1% for 300 nm (UV-B range)
 9      (Salby, 1996).
10           The combined effects of the earth's obliquity (the angle of the earth's axis of rotation with
11      respect to the plane of its orbit around the sun) and its precession (the rotation of the earth's axis
12      with respect to a perpendicular line  through the plane of its solar orbit) yield variations of up to
13      30% in total summertime solar flux, depending on latitude (Hartmann, 1994).
14
15      Zenith Angle: Latitude, Season and Time of Day
16           The sun's relative elevation is measured with respect to the vertical and is known as its
17      "zenith angle." This angle varies hourly, seasonally, and with latitude. Daily and seasonal
18      changes in solar zenith angle result in the largest changes in the magnitude of solar radiation
19      flux, with higher zenith angles corresponding to lower solar flux.  The largest natural fluxes
20      occur in the tropical regions, where solar noon occurs at a zenith angle at or near 0°. Seasonal
21      variation in solar flux ranges from small changes at the equator to very large changes at high
22      latitudes. Daily variations in solar flux, from sunrise to sunset, show added wavelength
23      dependence as a function of zenith angle, because transmission of some wavelengths are
24      sensitive to atmospheric pathlength due to scattering and absorption processes. These processes
25      will be discussed further below.
26
27      10.2.1.3  Atmospheric Radiative Interactions with Solar UV Radiation
28      Radiative Interactions in the Stratosphere
29           The stratosphere contains 90% or more of the total column density of O3, the principle gas
30      phase absorber of UV-B. Ozone interacts with UV radiation by scattering the photon or
31      absorbing and transforming its energy.  Upon absorbing a UV photon, O3 may photodissociate,
32      or become electronically- and vibrationally-excited.
        January 2005                              10-4       DRAFT-DO NOT QUOTE OR CITE

-------
 1           Photoabsorption by O3 occurs with very high efficiency. After electronically-excited O3
 2      (O3*) is formed, it will either dissociate into ground-state oxygen, O2, and an electronically
 3      excited oxygen radical, O(1D), or a collision between the O3 molecule with another gas molecule
 4      (M) may occur (See Reactions 1 and 2). Intermolecular collisions degrade the excess electronic
 5      energy of the O3 molecule by transferring it to other molecules as vibrational, rotational, and/or
 6      translational energies, which warm the atmosphere. The electronically excited oxygen radical
 7      can then react with H2O to form two hydroxyl radicals, OH (Reaction 3).  See Chapter 2 for
 8      further discussion of odd oxygen and HOX photochemistry.

                             O3 + hv -> O3* -> O^D) +  O2                               (10-1)
                                            -»O3*+M->O3 + M*                      (10-2)
 9
10
                                     O(1D) + H2O -» 2OH                              (10-3)
11
12
13      Either of these photochemical processes transforms the energy of the UV photon into a form
14      lacking the potential for human health or ecosystems  damage.
15           The WMO/UNEP (2002) scientific assessment reported that global average total column
16      O3 had declined by 3% from pre-1980 levels,  due to the presence of anthropogenic ozone-
17      depleting substances in the atmosphere. Ozone depletion has a strong latitude and seasonal
18      dependence.  The seasonality of total O3 changes differ between  the Northern and Southern
19      Hemispheres.  In the northern midlatitudes, total  column O3 declined by -4% during the
20      winter/spring seasons  and by approximately half that  amount in the summer/fall of the
21      1997-2001 time period, relative to pre-1980 total column O3 levels.  In southern midlatitudes,
22      total column O3 declined -6% during all seasons.
23           The concentration of O3 in a vertical column extending from the earth's surface is
24      expressed in Dob son Units (DU), corresponding to the column height in hundredths of a
25      millimeter of O3 at standard temperature and pressure (273 K and 1 atmosphere) (Wayne, 2000).
26      One DU = 2.587 x 1015 molecules of O3/cm2.  The total O3 column effectively prevents any
27      UV-C from reaching the surface, reduces the penetration of UV-B to the surface but does little to

        January 2005                              10-5       DRAFT-DO NOT QUOTE OR CITE

-------
1     attenuate the intensity of UV-A except at the shorter wavelengths close to the cut off for UV-B.
2     Cutchis (1974) calculated that with overhead sun a 10% decrease in the O3 column would lead to
3     20, 250, and 500% increases in flux at 305, 290, and 287 nm, respectively, values that have been
4     supported by ground observations in Toronto, ON (49° N; Kerr and McElroy, 1993). Rapid
5     changes of this magnitude appear to happen naturally. As seen in data collected by the Total
6     Ozone Mapping Satellite (TOMS) (Figure 10-2), the total O3 column undergoes wide natural
7     variation on short timescales (Cockell, 2001).
                                     Ozone Column (1990-1992)
               600
          c
          o
          (/)
          £1
          O
          a
                / / / / / /


                                                   Date
      Figure 10-2.  Ozone column abundances from the years 1990 to 1992 for 0, 40, and 80° N
                   as well as 80° S. The data for 80° S are incomplete, but the graph shows the
                   effects of the Antarctic O3 hole on total column abundances at this latitude.
                   The data for the Northern Hemisphere illustrate the natural variations in the
                   O3 column over time.  The data are taken from the TOMS (Total Ozone
                   Monitoring Satellite) data set (1979 to 1993).
      Source: Cockell (2001).
      January 2005
10-6
DRAFT-DO NOT QUOTE OR CITE

-------
 1           Nacreous and polar stratospheric clouds scatter radiation and aerosols, such as those
 2      injected into the stratosphere by explosive volcanic eruptions, both absorb and scatter radiation.
 3      Relative to the troposphere, the stratosphere is low in atmospheric pressure. Stratospheric clouds
 4      and aerosols are also more dispersed than in the troposphere.  Consequently, UV radiation can
 5      traverse the stratosphere with a substantially lower probability of encountering a gas molecule or
 6      cloud or aerosol particle than it would in the troposphere.  In the radiative transfer literature, the
 7      stratosphere is described as a "single scattering" regime for UV radiation, and UV that has
 8      penetrated the stratosphere is referred to as "direct beam UV." The troposphere, due to its high
 9      gas and particle concentrations is referred to as a "multiple scattering" regime.
10
11      Radiative Interactions in the Troposphere: Solar Irradiance vs. Actinic Flux
12           The troposphere contains < 10% of the total column O3 but -78% of the total atmospheric
13      mass including clouds, gas- and particle-phase radiation scatterers and absorbers, making it a
14      "multiple scattering" regime for UV radiation.  These scattering processes increase the mean-
15      free path a photon must travel before reaching the surface, transforming the direct beam UV
16      solar irradiance that has penetrated the stratosphere into diffuse, or actinic, UV (Briihl and
17      Crutzen, 1989).
18           Atmospheric scattering processes, such as Rayleigh and Mie scattering, are particle-size
19      dependent.  In Rayleigh scattering, gas molecules that are smaller than the wavelength of the
20      incident photon isotropically deflect incoming photons.  With Mie scattering, aerosol and cloud
21      droplets scatter incoming radiation with forward- and back-scattering tendencies. Actinic flux,
22      especially at the earth's surface, is directly  proportional  to surface albedo (Wendisch and Mayer,
23      2003). Surface albedo is very strongly wavelength dependent. For example, fresh and wet snow
24      reflects 60 to 90% of violet light, while soil and grass surfaces reflect < 5% of incident violet
25      light (Xenopoulos and Schindler, 2001). Wendisch and Mayer (2003) found, in their in situ
26      measurement and modeling study of the vertical distribution of solar irradiance,  that surface
27      albedos must be measured in order to accurately simulate solar flux, due to possibly large
28      variations in albedo within a given surface  type. Snow cover, even many kilometers from
29      measurement sites is known to increase detected UV irradiances. Complicated interactions
30      result when radiation is scattered by snow (or other bright surfaces) and backscattered or
31      absorbed by atmospheric particles and clouds in the same vicinity (WMO/UNEP, 2002).

        January 2005                               10-7        DRAFT-DO NOT QUOTE OR CITE

-------
 1      Variation in Solar Flux with Altitude
 2           Solar flux increases with altitude above sea level, due to the decreased presence of clouds
 3      and declining concentrations of scattering and absorbing atmospheric pollutants. Rayleigh
 4      scattering, discussed below, also lessens with decreasing atmospheric pressure. A number of
 5      measurements of UV radiation have been taken as a function of altitude and are reviewed by
 6      Xenopoulos and Schindler (2001).  Increases in flux as a function of altitude are given as percent
 7      irradiance enhancement per 1000 m relative to sea level. The effect can range from 9 to 24%
 8      /1000 m as function of the altitude at which the measurement was taken (Xenopoulos and
 9      Schindler, 2001). The effect corresponds to the relative pathlength traveled by the  solar photon:
10      flux is strongest when the photon is not impeded by atmospheric  scattering or absorbing agents.
11      Similarly, this effect is seen as a function of solar zenith angle, i.e., flux is at its maximum when
12      the atmospheric depth through which the photon must pass is at its shallowest.
13
14      Clouds
15           In principle, clouds have the largest influence on surface level UV irradiance, but their
16      effects are difficult to quantify. The depth and composition of a cloud determines,  in part, the
17      amount and wavelengths of radiation that it will scatter or absorb. Geometry is an  especially
18      important factor, as scattered or broken clouds may enhance, rather than reduce, surface flux.
19      For example, if the sun is not fully blocked by a cloud, the reduction in irradiance may be small
20      (WMO/UNEP, 2002). Quantifying the effect of clouds on surface UV flux, therefore, requires
21      information on cloud composition, geometry, altitude, and the position of the sun relative to the
22      cloud and the underlying surface, as a function of time. Provided that all of this detailed
23      information is available, a three dimensional model is then required to calculate surface-level
24      reductions or enhancements in UV flux.
25
26      Particulate Matter
27           On a zonally averaged basis, PM does not contribute significantly to lower tropospheric
28      absorption of UV radiation. However, in urban areas or other areas subject to high smog levels
29      (areas of significant biomass combustion),  PM may be the most important determinant of
30      ground-level erythemal UV flux, second only to cloud cover (U.S. Environmental Protection
31      Agency, 2004;  WMO/UNEP, 2003). Model-to-measurement comparisons of ground-level flux

        January 2005                               10-8        DRAFT-DO NOT QUOTE OR CITE

-------
 1      for Greece and Toronto have shown 20 and 5-10% reductions, respectively (McKenzie et al.,
 2      2001). Increases over the past 20 to 30 years in combustion-associated PM and black carbon
 3      may account for the inability to detect a surface trend in UV-B radiation caused by a known
 4      decrease in stratospheric O3 over the Northern Hemisphere (Barnard et al., 2003).
 5
 6      Gases
 1           In the upper troposphere, the UV-absorbing gases O3 and, of lesser importance,
 8      formaldehyde (CH2O) and SO2 are vented or diffuse from the surface. Stratospheric intrusions
 9      extrude O3-rich air into the troposphere where it mixes, increasing regional background O3 levels
10      (see Chapter 3). Tropospheric O3 data are typically expressed on a concentration basis, e.g.,
11      parts per billion by volume (ppbv), where 1 ppbv tropospheric O3 = 0.65 DU (IPCC, 2001).  The
12      mean values  of O3 in the free troposphere reported in the literature range from -50 to -80 ppbv,
13      with higher values occurring at the tropopause.  For example, a series of ozonesonde soundings
14      over France during the period from 1976 to 1995 showed an O3 increase from 48.9 ppbv in the
15      2.5 to 3.5 km layer to 56.5 ppbv in the 6.5 to 7.5 km layer, although the data revealed no
16      significant increasing trend over time (Ancellet and Beekmann, 1997).
17           Close to the surface in polluted urban settings, photochemistry produces a diurnal rise and
18      fall in O3 and PM concentrations. Temperature inversions tend to prevent the upward mixing
19      and dilution of ground-level O3, while trapping primary and secondary PM. A recent study of
20      the concurrence of O3 and PM is provided by Koloutsou-Vakakis et al.  (2001). No measurement
21      technique is currently available that can distinguish between absorption of incident UV radiation
22      by O3 versus absorption by PM.
23           UV absorption by gases becomes important under aerosol- and cloud-free conditions.
24      Figure 10-3 shows a calculation by Krotkov et al. (1998) of the sensitivity, as a function of
25      wavelength, of ground-level UV flux to a 1-DU decrease in total column O3 under cloud- and
26      aerosol-free conditions. A 1991 to 1992 study in Chicago in which ambient O3, broadband UV
27      irradiance and total sunlight were monitored (Frederick et al., 1993), produced a statistically
28      significant negative correlation between the UV irradiance and ambient O3 when the atmosphere
29      was relatively free of clouds and haze. Although they estimated that a 10-ppbv reduction in O3
30      was associated with a 1.3% increase in erythemally-weighted UV-B, they cautioned that this
31      figure had a comparatively large uncertainty (±1.2%).

        January 2005                             10-9        DRAFT-DO NOT QUOTE OR CITE

-------
                          290
                                        300            310            320
                                               Wavelength (nm)
                                                                                 330
       Figure 10-3.  The sensitivity of ground-level UV flux to a 1 DU change in total column O3,
                     under clear sky conditions, as a function of solar zenith angle (SZA).
       Source: Krotkov et al. (1998).
 1
 2
 3
 4
 5
 6
 7
 8
 9
10
11
12
10.2.1.4  Modeling Surface UV-B Irradiance
     The WMO/UNEP (2003) stated that, in principle, if the spatial distribution of all UV
absorbers and scatterers were fully known, the wavelength and angular distribution of the UV
irradiance at the earth's surface could be determined with model calculations.  However, the
poorly known and complicated distribution of the primary components (clouds, particles, O3,
and surface albedo) makes detailed predictions extremely difficult.
     Each of the factors described above must be integrated with human behaviors that also
influence potential exposure to UV-B radiation in order to calculate the human health risks
associated with small changes in total column O3 density. These human factors and the related
health consequences are described in detail below.
       January 2005
                                         10-10
DRAFT-DO NOT QUOTE OR CITE

-------
1      10.2.2  Factors Governing Human Exposure to Ultraviolet Radiation
2          Multiple, complex factors determine the solar flux of UV radiation at the ground level, as
3      discussed in Section 10.2.1. Figure 10-4 illustrates some of these geophysical/atmospheric
4      factors including stratospheric and tropospheric O3, clouds, aerosols, and Rayleigh scattering.
5      An estimate of the impact of changes in surface-level O3 on UV-related human health effects
6      requires information about human physical, behavioral, and demographic factors associated with
7      exposure to UV radiation.  In this section, the various factors that govern the degree of human
8      exposure to UV radiation are examined.
9
V
                      Backscattered
                       Radiation
                                      Incident Solar UV Radiation
                                        Stratospheric 03
      Figure 10-4.  Complexity of factors that determine human exposure to UV radiation.
                    In addition to the geophysical/atmospheric factors (e.g., stratospheric and
                    tropospheric O3, clouds, aerosols, and Rayleigh scattering) that affect the
                    solar flux of UV radiation at surface level, there are human physical,
                    behavioral and  demographic factors that influence human exposure to
                    UV radiation.
      January 2005
                                 10-11
DRAFT-DO NOT QUOTE OR CITE

-------
 1      10.2.2.1  Outdoor Activities
 2           Exposure to solar UV radiation is related to one predominating factor - time spent in
 3      outdoor microenvironments during daylight hours. A large U.S. study was conducted using the
 4      EPA National Human Activity Pattern Survey (NHAPS) to assess UV radiation dose in
 5      Americans (Godar, 2001; Godar et al., 2001, 2003). The EPA NHAPS recorded the activity
 6      profiles of 9,386 Americans (age 0 to 60+ years) over a 24-month period to assess their exposure
 7      to various environmental pollutants, plus UV radiation. This study indicated that there was a
 8      strong seasonal preference for outdoor activities, with people spending the most time outdoors
 9      during the summer, followed by spring, fall, and, lastly, winter (Godar et al., 2001).  Because the
10      solar erythemal (i.e., skin reddening) UV radiation dose is also highest during the summer, the
11      estimated UV radiation dose of Americans was more than ten-fold greater in the summer
12      compared to the winter season (Godar et al., 2001). Note that the error associated with
13      estimating UV radiation dose from exposure surveys and one EPA UV-monitoring site located at
14      each quadrant of the  U.S. will be high.
15           Vacationing at the beach during the summer was associated with higher UV radiation
16      exposures (Godar et  al., 2001; Thieden et al., 2001). Even after accounting for sunscreen use at
17      the beach, the erythemal UV radiation doses were more than 40% higher during a three-week
18      beach vacation compared to a three-week stay at home (Godar et al., 2001).  Danish children and
19      adolescents were found to receive > 50% of their annual UV radiation dose while vacationing at
20      European beaches  (Thieden et al., 2004a). The high UV radiation dose received at beaches is
21      due to increased time spent outdoors during daylight hours and increased risk behavior, namely
22      sunbathing. Sunbathing also was associated with increased annual UV radiation dose in the
23      Canadian National Survey on Sun Exposure and Protective Behaviours (Shoveller et al.,  1998).
24      Among the 3,449 adults (age 25+ years) who completed the telephone household survey, 21%
25      stated that they spent time actively sunbathing.  In  a Danish study with 164 participants, all
26      children (age 1 to 12 years) and teenagers (age 13 to 19 years) and 94% of adults (age 20 to
27      76 years) had days with risk behavior (Thieden et al., 2004b).  Teenagers had the highest annual
28      UV radiation doses, as compared to children and adults, a finding likely attributable to their
29      having the highest number of risk-behavior days.  Among teenagers, 76% of their UV radiation
30      dose during the measurement period was received on risk-behavior days, as determined using
31      personal electronic UV dosimeters and exposure diaries (Thieden et al., 2004b). In addition to

        January 2005                              10-12       DRAFT-DO NOT QUOTE OR CITE

-------
 1      vacationing and sunbathing, participation in outdoor sports (e.g., basketball, soccer, golfing,
 2      swimming, cycling) also significantly increased UV radiation exposure (Moehrle, 2001; Moehrle
 3      et al., 2000; Moise et al., 1999; Thieden et al., 2004a,b).
 4
 5      10.2.2.2 Occupation
 6           Of the various factors that affect human exposure to UV radiation, occupation is also
 7      important.  Approximately 5% of the American workforce perform work in outdoor
 8      microenvironments, as determined by the EPA NHAPS (Godar et al., 2001).  American indoor
 9      workers, on average, spend -10% of their day outdoors and are exposed to -30% of the total
10      ground-level UV flux, as measured by the EPA UV-monitoring program, during this time period
11      (Godar et al., 2001).  In comparison to indoor or home workers, outdoor workers are exposed to
12      much higher levels of UV radiation (Kimlin et  al., 1998a; Thieden et al., 2004a), frequently at
13      levels that are above current exposure limits set by the International Commission on
14      Non-Ionizing Radiation Protection (ICNIRP, 2004). For example, Thieden et al. (2004a)
15      observed that the annual UV radiation dose, estimated using personal electronic UV dosimeters
16      and exposure diaries, -70% higher for gardeners than indoor workers. The gardeners received
17      the majority (55%) of their UV radiation dose on working days (Thieden et al.,  2004a). Another
18      study found that outdoor workers received three to four times the annual UV radiation exposure
19      of indoor workers (Diffey, 1990). At-risk working populations include farmers (Airey et al.,
20      1997; Schenker et al., 2002), fisherman (Rosenthal et al., 1988), landscapers (Rosenthal et al.,
21      1988), building and construction workers (Gies and Wright, 2003), physical education teachers
22      (Vishvakarman et al., 2001), mail delivery personnel (Vishvakarman et al., 2001), and various
23      other workers who spend the majority of their day in outdoor microenvironments during peak
24      UV radiation hours.
25
26      10.2.2.3 Age
27           Age-related differences in UV radiation dose was examined in a U.S. study using the EPA
28      NHAPS (Godar, 2001; Godar et al., 2001).  The average UV radiation dose among American
29      children (age < 12 years) was  estimated to be similar to that of adults (age 20+ years), but -20%
30      higher than that of adolescents (age 13 to 19 years) (Godar, 2001). During the summer, higher
31      UV radiation doses were observed in children < 5 years of age and men 40+ years of age (Goder,

        January 2005                             10-13       DRAFT-DO NOT QUOTE OR CITE

-------
 1      2001).  In a large Canadian survey, 89% of children (age < 12 years) had 30 minutes or more of
 2      daily UV exposure, compared to 51% for both adults (age 25+ years) and youth (age 15 to
 3      24 years) (Lovato et al., 1998a,  1998b; Shoveller et al., 1998).  In an English study (Diffey et al.,
 4      1996), UV radiation exposure was estimated in 180 children (age 9 to 10 years) and adolescents
 5      (age 14 to 15 years) using personal film badges and exposure records.  Once again, children were
 6      found to have received higher UV radiation exposure compared to adolescents (Diffey et al.,
 7      1996).  However, in contrast to the results from the studies discussed above, a Danish study
 8      observed that the annual UV radiation dose in teenagers (age 13 to  19 years) was -20% higher
 9      compared to children (age 1 to 12 years) (Thieden  et al., 2004b).  This increase in UV radiation
10      dose in teenagers was attributed to their increased risk-behavior days.  The time profiles of daily
11      UV radiation exposure among grade 8 students was assessed in an Australian study using a
12      routinely operating UV-Biometer and questionnaires (Moise et al.,  1999). The results indicated
13      that up to 47% of the daily UV radiation dose fell within the time periods when students were
14      outdoors during school hours, sitting under shaded structures during lunch breaks and
15      participating in routine outdoors or sports activities (Moise et al., 1999).
16           Two studies examined lifetime UV radiation  exposure among persons in the U.S. (Godar
17      et al., 2001) and Denmark (Thieden et al., 2004b).  Both studies observed that while there are
18      slight differences in UV radiation dose by age, generally people receive fairly consistent UV
19      doses at different age intervals throughout their lives.
20
21      10.2.2.4 Gender
22           Studies have indicated that females generally spend less time  outdoors and, consequently,
23      have lower UV radiation exposure compared to males (Gies et al., 1998;  Godar et al., 2001;
24      Shoveller et al.,  1998). The U.S. study by Godar et al. (2001) observed that even though both
25      males and females had somewhat consistent erythemal UV radiation doses throughout their
26      lives, males received consistently higher UV doses compared to females  at all age groups.
27      Among all Americans, the lowest exposure to UV radiation was received in females during their
28      child-raising years (age 22 to 40 years) (Godar et al., 2001).  The highest exposure was observed
29      in males aged 41 to 59 years in the U.S. study (Godar et al., 2001), while a similar Canadian
30      survey found that younger adult males had the greatest exposures to UV radiation (Shoveller
31      etal., 1998).

        January 2005                              10-14       DRAFT-DO NOT QUOTE OR CITE

-------
 1      10.2.2.5  Geography
 2           In the U.S. study by Godar et al. (2001), erythemal UV radiation doses were examined in
 3      persons living in northern and southern regions.  Northerners and southerners were found to
 4      spend an equal amount of time outdoors; however, the higher solar flux at lower latitudes
 5      significantly increased the annual UV radiation dose for southerners (Godar et al., 2001). The
 6      annual UV radiation doses in southerners were 25 and 40% higher in females and males,
 7      respectively, compared to northerners (Godar et al., 2001). Other studies also have shown that
 8      altitude and latitude influence personal exposure to UV radiation (Kimlin et al., 1998b; Rigel
 9      etal., 1999).
10
11      10.2.2.6  Protective Behavior
12           Protective behaviors such as using sunscreen (e.g., Nole and Johnson, 2004), wearing
13      protective clothing (e.g., Rosenthal et al., 1988; Sarkar, 2004; Wong et al., 1996), and spending
14      time in microenvironments with shaded surfaces (Moise et al., 1999; Parisi et al., 1999) have
15      been shown to reduce exposure to UV radiation. In one study, the use of sunscreen was
16      associated with extended intentional UV radiation exposure (Autier et al.,  1999); however, a
17      follow-up study indicated that sunscreen use increased duration of exposures to doses of UV
18      radiation that were below the threshold level for erythema (Autier et al., 2000).
19           In a national study of U.S. youths aged  11 to 18 years, the most prevalent protective
20      behavior was sunscreen use (39.2%) followed by use of a baseball hat (4.5%) (Davis et al.,
21      2002).  There were significant differences in the use of sunscreen by age group and gender, with
22      the younger age group (age 11 to 13 years) and girls having greater likelihood (47.4 and  48.4%,
23      respectively) of using sunscreen (Davis et al., 2002). The Canadian National  Survey on  Sun
24      Exposure and Protective Behaviours observed that less than half of the adults (age 25+ years,
25      n = 3,449) surveyed took adequate protective actions (Shoveller et al., 1998).  Once again,
26      children (age < 12 years,  n = 1,051) were most protected from exposure to UV radiation, with
27      76% using sunscreen and 36% avoiding the sun, as reported by their parents (Lovato et al.,
28      1998a). However, the protection level was still not adequate, as indicated by  the high 45% rate
29      of erythema in children. Among Canadian youth (age 15 to 24 years,  n = 574), protective
30      actions from UV radiation exposure included wearing a hat (38%) and seeking shade and
31      avoiding the sun between the peak hours of 11:00 a.m. to 4:00 p.m. (26%) (Lovato et al., 1998b).

        January 2005                              10-15       DRAFT-DO NOT QUOTE OR CITE

-------
 1      The lowest prevalence of protective behavior among the youth was likely responsible for the
 2      highest proportion of erythema (68%) experienced in this age group. A Danish study observed
 3      that both children and teenagers applied sunscreen on more days than adults, but teenagers had
 4      the most days with erythema, due to their increased risk behavior (Thieden et al., 2004b).
 5      A survey in Switzerland of 1,285 individuals, including children and parents, indicated that
 6      sunscreen use was the protective action most commonly used, but only at the beach and not in
 7      routine daily exposure (Berret et al., 2002).  Protective clothing and avoiding the sun were not
 8      highly used among these individuals.
 9
10      10.2.2.7 Summary of Factors that  Affect Human Exposures to Ultraviolet Radiation
11           The factors that potentially influence UV radiation doses were discussed in the previous
12      sections and include choice of leisure activities, occupation, age, gender, geography, and
13      protective behavior.  Results from the various studies indicate that specific populations may be at
14      risk for higher exposures to UV radiation. Of particular concern are the following potentially
15      susceptible populations:
16          •  Individuals who engage in high-risk behavior, viz., sunbathing;
17          •  Individuals who participate in outdoor sports and activities, including professional
               athletes;
18          •  Individuals who work in outdoor microenvironments with inadequate shade, e.g.,
               farmers, fishermen, gardeners, landscapers, building and construction workers;
19          •  Young children; and
20          •  Individuals living in geographic areas with higher solar flux (i.e., lower latitudes
               [e.g., Honolulu, HI] and higher altitudes [e.g.,  Denver, CO]).
21
22      10.2.3  Factors Governing Human Health Effects due to Ultraviolet Radiation
23           Ultraviolet radiation occupies a specific region of the  electromagnetic spectrum of
24      wavelengths and can be further subdivided into three parts,  UV-A (320 to 400 nm), UV-B
25      (280 to 320 nm), and UV-C (200 to 280 nm). Most of the health risks associated with UV
26      radiation exposure are wavelength dependent.  Wavelengths < 180 nm are of little practical
27      biological significance as they are readily absorbed in the air (ICNIRP, 2004). Until 1980,
28      it was generally thought that wavelengths <  315 nm were responsible for the most significant

        January 2005                             10-16       DRAFT-DO NOT QUOTE OR CITE

-------
 1      adverse UV radiation health effects; however, recent studies have found that the longer
 2      wavelengths in the UV-A range also might produce adverse responses at substantially higher
 3      doses (ICNIRP, 2004).
 4           Action spectra of a given biological response to UV radiation across its spectral range are
 5      used to estimate exposure by weighting individual wavelength intensities according to the
 6      associated response. The overall effectiveness of the incident flux at inducing the biological
 7      response of interest is computed by means of the relationship
 8
                                  effective if radiance = )I^E^a"X                           (10-4)
                                                       A
 9
10      where 1A and EA are, respectively, the irradiance and its relative effectiveness at wavelength A.
1 1           UV-A and UV-B radiations differ in their abilities to initiate DNA damage, cell signaling
12      pathways, and immune alterations.  In this section, the various adverse health effects associated
13      with acute and chronic UV radiation exposure will be discussed.
14
15      10.2.3.1  Erythema
1 6      Association Between Ultraviolet Radiation Exposure and Erythema
17           The most conspicuous and well-recognized acute response to UV radiation is erythema, or
18      the reddening of the skin, which is likely caused by direct damage to DNA by UV-B and UV-A
19      radiation (Matsumura and Ananthaswamy, 2004). Indirect oxidative damage also might occur at
20      longer wavelengths (Matsumura and Ananthaswamy, 2004). Skin type appears to play a large
21      role in the sensitivity  to UV radiation -induced erythema.  The Fitzpatrick classifications for skin
22      types are: (1) skin type I - individuals with extremely sensitive skin that sunburns easily and
23      severely, and is not likely to tan (e.g., very fair skin, blue eyes, freckles); (2) skin type
24      II - individuals with very sensitive skin that usually sunburns easily and severely, and tans
25      minimally (e.g., fair skin, red or blond hair, blue, hazel or brown eyes); (3) skin type
26      III - individuals with sensitive skin that sunburns moderately and tans slowly (e.g., white skin,
27      dark hair); (4) skin type IV - individuals  with moderately sensitive skin that sunburns minimally
28      and usually tans well  (e.g., white or light brown skin, dark hair, dark eyes); (5) skin type
29      V - individuals with minimally sensitive  skin that rarely sunburns and tans deeply (e.g., brown
        January 2005                              10-17       DRAFT-DO NOT QUOTE OR CITE

-------
 1      skin); and (6) skin type VI - individuals with nonsensitive skin that never sunburns and tans
 2      profusely (e.g., dark skin). A study by Harrison and Young (2002) found that the perceptible
 3      minimal erythemal dose was approximately two-fold greater among individuals with skin type
 4      IV compared to skin type I, although there was considerable overlap in the minimal erythemal
 5      dose among the four skin types. Waterston et al. (2004) further observed that within an
 6      individual, erythemal response differed by body site (e.g., abdomen, chest, front upper arm, back
 7      of thigh). These differences were likely attributable to body site-specific variations in melanin
 8      pigmentation.
 9          Kollias et al. (2001) investigated the change in erythemal response following a previous
10      exposure to UV radiation. Body sites that received a second exposure to UV radiation always
11      showed a reduced erythemal response compared to body sites with a single exposure, especially
12      when the first exposure was at levels greater than the minimal erythemal dose. The suppression
13      of erythema was more pronounced when the second exposure was given 48 hours after the first.
14      However, Kaidbey and Kligman (1981) found that multiple exposures to subthreshold doses of
15      UV  radiation at 24-hour intervals were found to lower the minimal erythemal dose.  Henrisken
16      et al. (2004) observed that the change in threshold depended on skin type. In 49 healthy
17      volunteers with skin types II, III, and IV, just perceptible erythema 24 hours post-exposure was
18      chosen as the minimal erythemal dose. After four days of repeated UV radiation, the minimal
19      erythemal dose was lowered by 40 to 50% in darker-skinned persons.  However, among fair-
20      skinned individuals, there was no change in the erythemal threshold dose with repeated exposure
21      to UV radiation.
22          A reference erythema action spectrum was adopted by the Commission Internationale de
23      1'Eclairage (International Commission on Illumination, CIE) in 1987 (McKinlay and Diffey,
24      1987). The CIE erythema action spectrum indicates that UV-B radiation is orders of magnitude
25      more effective per unit dose than UV-A radiation.  However, a follow-up study by Diffey (1994)
26      that  compared the observed and predicted minimal erythemal doses found that the erythemal
27      sensitivity of skin at longer UV wavelengths (> 350 nm) was greater than predicted from the CIE
28      reference action spectrum.
29
30
        January 2005                             10-18       DRAFT-DO NOT QUOTE OR CITE

-------
 1      Risk of Erythema from Changes in Tropospheric Ozone Levels
 2           There is no literature on the risk of erythema associated with changes in tropospheric or
 3      ground-level O3 levels. However, one study conducted a risk assessment of the effects of
 4      stratospheric O3 depletion on the risk of erythema (Longstreth et al., 1998). Stratospheric O3
 5      depletion will result in only a slight increase influence rate of UV-A as O3 only absorbs a very
 6      little part of this UV spectrum. However, the ground-level UV-B flux will likely increase as O3
 7      absorbs radiation in that wavelength range with high efficiency.
 8           The risk analysis by Longstreth et al. (1998) concluded that erythema will not appreciably
 9      increase with depletion of the O3 layer.  This is due to the powerful adaptation of the skin to
10      different levels of UV radiation, as evidenced by its ability to cope with changes in UV radiation
11      by season (van der Leun and de Gruijl, 1993).  Gradual exposure to increasing UV radiation
12      from the winter to summer leads to decreased sensitivity of the skin. In midlatitudes, the UV-B
13      radiation in the summer is ten-fold greater than in the winter. In contrast, the steady depletion of
14      the O3 layer has been estimated to likely lead to an -20% increase in UV-B in 10 years
15      (Longstreth et al., 1998).  Such a comparatively small increase in UV radiation throughout the
16      years, therefore, would not seem likely to significantly increase the risk of erythema. Given that
17      the stratospheric O3 depletion is estimated as not being likely to affect the risk of erythema, one
18      can conclude that changes in ground-level O3 (which only constitutes no more than 10% of total
19      atmospheric O3) is not likely to result in increased risk.
20
21      10.2.3.2  Skin Cancer
22           According to the Skin Cancer Foundation, one in six Americans will develop skin cancer
23      during their lifetime (Gloster and Brodland, 1996).  The three main forms of skin cancer include
24      basal cell carcinoma and squamous cell carcinoma, which are both nonmelanoma skin cancers,
25      and malignant melanoma. Nonmelanoma skin cancers constitute  more than one-third of all
26      cancers in the U.S. and -90% of all skin cancers, with basal cell carcinoma being approximately
27      four times as common  as  squamous cell carcinoma (Diepgen and Mahler, 2002; ICNTRP, 2004).
28      The incidence of malignant melanoma (3 to 4% of all cancers) is much lower compared to
29      nonmelanoma skin cancers, but melanoma has great metastatic potential and accounts for the
30      majority of skin cancer deaths (Jemal et al., 2004).
        January 2005                              10-19       DRAFT-DO NOT QUOTE OR CITE

-------
 1           Exposure to UV radiation is considered to be a major risk factor for all three forms of skin
 2      cancer (Gloster and Brodland,1996; Diepgen and Mahler, 2002; IARC,  1992). Ultraviolet
 3      radiation is especially effective in inducing genetic mutations and acts as both a tumor initiator
 4      and promoter. Keratinocytes have evolved DNA repair mechanisms that correct the damage
 5      induced by UV, but mutations  do occur, leading to skin cancers that are appearing with
 6      increasing frequency (Hildesheim and Fornace, 2004). The relationship between skin cancer and
 7      chronic exposure to UV radiation is further explored below, followed by discussion of the
 8      influence of O3 on the incidence of skin cancer.
 9
10      10.2.3.3   Ultraviolet Radiation Exposure and the Incidence of Nonmelanoma
11                Skin Cancers
12           The incidence of all three types of cancers has been shown to rise  with increasing UV
13      radiation concentrations across the U.S. (de Gruijl, 1999);  however, the most convincing
14      evidence for a causal relationship exists between UV radiation and squamous cell carcinoma.
15      Squamous cell carcinoma occurs almost exclusively  on skin that is regularly exposed to the sun,
16      such as the face, neck,  arms, and hands. The incidence is higher among whites in areas of lower
17      latitudes, where solar flux is greater (Kricker et al., 1994). The risk of squamous cell carcinoma
18      was shown to increase with life-long accumulated exposure to UV radiation in one cross-
19      sectional study (Vitasa et al., 1990), but was found to be associated only with the exposure ten
20      years prior to diagnosis in a case-control study (Gallagher et al., 1995a). One of the major
21      concerns with both types of studies is the potential for recall bias in reporting their UV radiation
22      exposure as individuals are already aware of their disease status.
23           Ultraviolet radiation also has been linked to basal cell carcinoma.  Basal cell carcinoma is
24      common on the face and neck (80-90%) but rarely occurs on the back of the hands (de Gruijl,
25      1999).  While cumulative UV radiation exposure was not associated with risk of basal cell
26      carcinoma (Vitasa et al., 1990), increased risk was observed in individuals with increased
27      recreational UV radiation exposure in adolescence and childhood (age < 19 years) and
28      individuals with a history of severe erythema in childhood (Gallagher et al., 1995b).  Once again,
29      consideration must be given to potential recall bias in assessing these results.   Thus, there is
30      suggestive evidence that UV radiation also plays a role in the development of basal cell
31      carcinoma, but the etiologic  mechanisms for squamous cell carcinoma and basal cell carcinoma
32      likely differ.  In an Australian study conducted in a subtropical community, having fair skin,

        January 2005                              10-20       DRAFT-DO NOT QUOTE OR CITE

-------
 1      a history of repeated sunburns, and nonmalignant solar skin damage diagnosed by dermatologists
 2      were strongly associated with both types of nonmelanoma skin cancer (Green et al., 1996).
 3      Outdoor occupation was not associated with nonmelanoma skin cancer, which was likely due to
 4      significant self-selection. Individuals with fair or medium complexions and a tendency to
 5      sunburn, though they accounted for more than 80% of the community study sample, were
 6      systematically underrepresented among outdoor workers (Green et al.,  1996). This selection bias
 7      may partly explain the lack of consistent quantitative evidence of a causal link between UV
 8      radiation and skin cancer in humans.
 9           De Gruijl et al. (1993) assessed the action spectrum for nonmelanoma skin cancers using
10      hairless albino mice. Human data are not available to examine the wavelength dependence of
11      the carcinogenicity of UV radiation. After adjusting for species differences, the Skin Cancer
12      Utrecht-Philadelphia action spectrum indicated the highest effectiveness in the UV-B range with
13      a maximum at 293 nm, which dropped to 10~4 of this maximum at the UV-A range above
14      340 nm (de Gruijl et al., 1993).  The mutations commonly present in thep53 tumor suppressor
15      gene in individuals with squamous cell carcinoma and basal cell carcinoma are called the
16      "signature" mutations of UV-B radiation (de Gruijl, 2002).  UV-B radiation is highly mutagenic
17      due to the fact that DNA is  a chromophore for UV-B, but not for UV-A, radiation (Ichihashi
18      et al., 2003). However, other studies have found that UV-A radiation, in  addition to UV-B
19      radiation, can induce DNA  damage (Persson et al, 2002; Ruenger et al., 2000). DNA damage by
20      UV-A is mediated by reactive oxygen species, thus is indistinguishable from damage caused by
21      other agents that generate reactive oxygen species (de Gruijl, 2002).  Epidemiologic evidence of
22      a carcinogenic effect of UV-A was found in a study of psoriasis patients receiving oral psoralen
23      and UV-A radiation treatment (Stern et al., 1998). High-dose exposure to oral psoralen and
24      UV-A radiation was associated with a persistent, dose-related increase in the risk of squamous
25      cell cancer, but with much less effect on the risk of basal cell cancer. Therefore, although UV-B
26      radiation has long been considered the main culprit for nonmelanoma skin cancer, limited
27      evidence suggest that UV-A radiation may also play a role.
28           Susceptible populations for nonmelanoma skin cancers include individuals with reduced
29      capacity for nucleotide excision repair, the primary repair mechanism for UV radiation-induced
30      DNA lesions (Ichihashi et al., 2003).  At particular risk are individuals with xeroderma
31      pigmentosum, as they have defective nucleotide excision repair in all tissues (Kraemer, 1997;

        January 2005                              10-21       DRAFT-DO NOT QUOTE OR CITE

-------
 1      Sarasin, 1999). Skin type also largely affects susceptibility to skin cancer. Of the six skin
 2      phenotypes, the most sensitive individuals are those with skin types I and II, who have a fair
 3      complexion, blue or green eyes, and red or blond hair (Diepgen and Mahler, 2002).  These
 4      individuals tend to sunburn easily, tan poorly, and freckle with sun exposure. A history of
 5      repeated sunburns also appears to increase the risk of both cancers, while sunburns during
 6      childhood are more associated with increased basal cell carcinoma (Gallagher et al., 1995b;
 7      Green etal., 1996).
 8
 9      Ultraviolet Radiation and the Incidence of Cutaneous Malignant Melanoma
10           From 1973 to 1994, the incidence rate of melanoma increased 120.5% along with an
11      increased mortality rate of 38.9% among whites in the U.S. (Hall et al., 1999). The ICNIRP
12      (2004) states that during the past 40 years or so, each decade has seen a two-fold increase in the
13      incidence of malignant melanoma in white populations, with increased incidence observed more
14      prominently in individuals living in lower latitudes.  Cutaneous malignant melanoma has
15      mutifactorial etiology, with environmental, genetic, and host factors (Lens and Dawes, 2004).
16      While the major environmental factor of melanoma has been identified as UV radiation
17      exposure, the risk of melanoma appears to depend on the interaction between the nature of the
18      exposure and skin type (Lens  and Dawes, 2004).
19           Fears et al. (2002) examined the association between invasive cutaneous melanoma and
20      UV radiation in non-Hispanic whites using a case-control study design. Lifetime residential
21      history was coupled with mid-range UV-B radiation flux measurements to reduce exposure
22      misclassification and recall bias.  A 10% increase in the average annual UV-B flux was
23      significantly associated with a 19% increase in individual odds for melanoma in men and a
24      16% increase in women. Whiteman et al. (2001) conducted a systematic review of studies that
25      examined the association between childhood UV radiation exposure and risk of melanoma.
26      Researchers found that ecological studies assessing ambient sun exposure consistently reported
27      higher risks of melanoma among people who resided in an environment with high UV radiation
28      during their childhood (Whiteman et al., 2001).  The lack of consistency among the case-control
29      studies was likely due to the varying methods used to assess UV radiation dose.
30           While the evidence is generally suggestive of a causal relationship between UV radiation
31      and malignant melanoma, possibly conflicting data has been observed. For example, the highest

        January 2005                              10-22       DRAFT-DO NOT QUOTE OR CITE

-------
 1      occurrence of malignant melanoma is on men's backs and women's legs, areas that do not have
 2      prolonged exposure to the sun (Rivers, 2004). This indicates that malignant melanoma tends to
 3      occur in sites of intermittent, intense sun exposure (trunk and legs), rather than in areas of
 4      cumulative sun damage (head, neck, and arms), in contrast to nonmelanoma skin cancers
 5      (Swetter, 2003).
 6          The available data conflict with regard to the relative importance of UV-A versus UV-B in
 7      inducing melanomas. UV-A has a much higher flux rate at the earth's surface, as it is not
 8      absorbed by O3 and it is able to penetrate more deeply into the skin surface  due to its longer
 9      wavelength. However, UV-B, as mentioned earlier, is much more energetic and, therefore, more
10      effective in photochemically altering DNA. The individual roles of UV-B and UV-A in the
11      development of cutaneous malignant melanoma have been examined in several studies.
12      A case-control study of 571 patients and 913 matched controls observed an  elevated odds ratio
13      for development of malignant melanoma in individuals who regularly used  tanning beds, which
14      typically are UV-A sources (Westerdahl et al., 2000).  In a study by Setlow et al. (1993),
15      an action spectrum using the tropical fish Xiphophorus indicated that UV-A range wavelengths
16      were especially important in malignant melanoma induction. However, an  action spectrum
17      using the opossum Monodelphis domestica found that the potency of UV-A for melanoma
18      induction was extremely low compared to that of UV-B (Robinson et al., 2000). A recent study
19      by De Fabo et al. (2004) examined the differences in wavelength effectiveness using a
20      hepatocyte growth factor/scatter factor-transgenic mouse model. The epidermal tissue of these
21      transgenic mice behaves very similarly to the human epidermis in response  to UV exposure.
22      Given the absence of a mammalian melanoma action spectrum, the standardized CIE erythema
23      action spectrum was used to deliver identical erythemally effective doses. Only UV-B radiation
24      was found to initiate mammalian cutaneous malignant melanoma.  UV-A radiation, even at
25      doses considered physiologically relevant, were ineffective at inducing melanoma (De Fabo
26      et al., 2004).  Overall, current evidence suggests that UV-B, and not UV-A, is the primary risk
27      factor for malignant melanoma (ICNIRP, 2004).
28          The populations susceptible for malignant melanoma are similar to those for nonmelanoma
29      skin cancers. Once again, individuals with xeroderma pigmentosum or a reduced capacity of
30      nucleotide excision repair are at increased risk (Tomescu et al., 2001; Wei et al., 2003).
31      Individuals with skin types I and II, or the fair-skin phenotype (blue or green eyes; blond or red

        January 2005                             10-23       DRAFT-DO NOT QUOTE OR CITE

-------
 1      hair; skin that freckles, sunburns easily, and does not tan), have increased susceptibility to
 2      malignant melanoma (Evans et al., 1988; Swetter, 2003; Veier0d et al., 2003). However, the
 3      incidence of melanoma was also positively associated with UV radiation in Hispanics and blacks
 4      (Hu et al., 2004). Although the incidence of melanoma is much lower in Hispanics and blacks
 5      compared to whites, melanomas in these populations are more likely to metastasize and have a
 6      poorer prognosis (Black et al., 1987; Bellows et al., 2001).  Among children, malignant
 7      melanoma appears to have similar epidemiologic characteristics to the adult form of the disease
 8      (Whiteman  et al., 1997).  Individuals with intermittent, intense sun exposure, particularly during
 9      childhood, were found to have increased risk of melanoma (Whiteman et al., 2001), in contrast
10      to the association between cumulative exposure and increased risk of squamous cell carcinoma.
11      One study found that a personal history of nonmelanoma skin cancer or precancer, higher
12      socioeconomic status, and increased numbers of nevocytic nevi also were associated with
13      increased incidence of melanoma (Evans et al., 1988).
14
15      Effect of Changes in Tropospheric Ozone Levels on Skin Cancer Incidence
16           The current evidence strongly suggests a causal link between exposure to UV radiation and
17      the incidence of both nonmelanoma and melanoma  skin cancer. Genetic factors, including skin
18      phenotype and ability to repair DNA, affect an individual's susceptibility to skin cancer.
19      Quantifying the relationship between UV radiation  and skin cancer is complicated by  the
20      uncertainties involved in the selection of an action spectrum and appropriate characterization of
21      dose (e.g., peak or cumulative levels of exposure, childhood or lifetime exposures). In addition,
22      there are multiple complexities in attempting to quantify the effect of tropospheric O3  levels on
23      UV-radiation exposure, as described in Section 9.11.2. There is an absence of published  studies
24      that critically examine  any increased incidence of skin cancer that may be attributable to
25      decreased tropospheric O3 exposures (which reflects the significant challenges in determining
26      ground-level O3-related changes in UV radiation exposure).
27           Several studies have examined the potential effect of stratospheric O3 depletion on the
28      incidence of skin cancer (de Gruijl, 1995; Longstreth et al., 1995; Madronich and de Gruijl,
29      1993; Slaper et al., 1996). Stratospheric O3 depletion is likely to increase the ground-level UV-B
30      flux, as O3 absorbs radiation in that wavelength range with high efficiency.  Because UV-B
31      radiation is  primarily implicated in the induction of skin cancer, especially among persons with

        January 2005                              10-24      DRAFT-DO NOT QUOTE OR CITE

-------
 1      skin phenotypes I and II, there is concern that the depletion of the O3 layer will result in
 2      significantly increased incidence of skin cancers.
 3           Estimation of the increased risk in melanoma associated with O3 depletion cannot be done
 4      adequately due to the lack of a mammalian action spectrum for melanoma. In addition, the
 5      complexity of the UV-related induction mechanism of melanoma adds an additional layer of
 6      uncertainty to the calculations.  The excess risk in nonmelanoma skin cancers associated with a
 7      decrease in stratospheric O3 was estimated using the Skin Cancer Utrecht-Philadelphia action
 8      spectrum based on hairless albino mice (Longstreth et al., 1995). Quantification of how much
 9      more UV radiation would reach ground level with each percentage decrease in O3 required
10      several assumptions:  (1) annual doses  are an appropriate measure; (2) personal doses  are
11      proportional to ambient doses; and, most notably, (3) each percentage decrease in O3 is
12      associated with a 1.2% increase in UV radiation. Next, the relationship between UV radiation
13      and nonmelanoma skin cancer incidence was determined.  Each percent increase in annual UV
14      radiation dose was estimated to cause a 2.5% increase in squamous cell carcinoma and 1.4%
15      increase in basal cell carcinoma over a human lifetime. Incorporating all these factors,
16      Longstreth et al. (1995) calculated that a sustained 10% decrease in stratospheric O3
17      concentration would result in 250,000 additional nonmelanoma skin cancer cases per year.
18      Madronich and de Gruijl (1993) noted that the largest percent of O3-induced nonmelanoma skin
19      cancer increases would be at high latitudes, where baseline incidence of skin cancer is usually
20      small. Assuming a phaseout of primary O3-depleting substances by 1996, as established by the
21      Copenhagen Amendments in 1992, Slaper et al. (1996) estimated that the number of excess
22      nonmelanoma skin cancers in the U.S.  caused by O3 depletion would exceed 33,000 per year
23      (or ~7 per 100,000) around the year 2050.
24           However, estimating the increase in nonmelanoma skin cancer incidence attributable to the
25      depletion of the stratospheric O3 layer is marred by uncertainty. The following statement by
26      Madronich and de Gruijl (1994) describes the uncertainty of estimating the effect of
27      stratospheric O3 depletion on the incidence of skin cancer:
z9                Extrapolating trends and effects of UV into the future is very hypothetical due to
30                uncertainties that arise from atmospheric chemistry, epidemiology, and related
31                disciplines. The values that we calculated are one plausible measure of the
32                magnitude of the 03-UV effects... .The timescales for atmospheric change and skin-
33                cancer development are still far from certain: 03 reductions are expected to continue
34                well into next century, and the time between UV exposure and development of skin
35                cancer is essentially unknown. . .

        January 2005                             10-25       DRAFT-DO NOT QUOTE OR CITE

-------
 1      Therefore, much caution is necessary when assessing and interpreting the quantitative results of
 2      excess nonmelanoma skin cancer incidence due to stratospheric O3 depletion.  Note that although
 3      the effect of changes in ground-level or tropospheric O3 concentrations on skin cancer incidence
 4      has not been assessed, it would be expected to be much less compared to the effect from the
 5      depletion of the stratospheric O3 layer, given that tropospheric O3 makes up <  10% of the total
 6      atmospheric O3.
 7
 8      10.2.3.4   Ocular Effects of Ultraviolet Radiation Exposure
 9      Ultraviolet Radiation Exposure and Risk of Ocular Damage
10           Ocular damage from UV radiation exposures includes effects on the cornea, lens, iris, and
11      associated epithelial and conjunctival tissues. Absorption of UV radiation differs by
12      wavelength, with short wavelengths (< 300 nm) being almost completely absorbed by the
13      cornea, whereas longer wavelengths are transmitted through the cornea and absorbed by the lens
14      (McCarty and Taylor, 2002).  The most common acute ocular effect of environmental UV
15      exposure is photokeratitis, also known as snowblindness, caused by absorption of short
16      wavelength UV radiation by the cornea. The action spectrum indicated that maximum
17      sensitivity of the human eye was found to occur at 270 nm (ICNIRP, 2004; Pitts, 1993). The
18      threshold for photokeratitis  in humans varied from 4 to 14 mJ/cm2 for wavelengths 220 to
19      310 nm.  For UV-A radiation, levels exceeding 10 J/cm2 were necessary for corneal injury.
20           Exposure to longer wavelengths has been shown to cause both transient and permanent
21      opacities of the lens, or cataracts.  There is extensive toxicologic and epidemiologic evidence
22      supporting the causal association between UV radiation and cataracts (Hockwin et al., 1999;
23      McCarty and Taylor, 2002). Ultraviolet radiation-induced cataracts are hypothesized to be
24      caused by oxidative stress leading to increased reactive species in the lens, which then causes
25      damage to lens DNA and cross-linking of proteins. Exposure time to low-dose UV radiation was
26      found to strongly influence  cataract formation (Ayala et al., 2000).  An action spectrum
27      determined using young female rats indicated that the rat lens was most sensitive to 300 nm,
28      correcting for corneal transmittance (Merriam et al., 2000). Oriowo et al. (2001) examined the
29      action spectrum for cataract formation using whole cultured lens from pigs. As pigs lens are
30      similar in shape and size to  the human lens, some inferences may be made.  Results indicated
31      that the 270 to 315 nm waveband was most effective in producing UV-induced cataracts in vitro.

        January 2005                              10-26       DRAFT-DO NOT QUOTE OR CITE

-------
 1      However, the threshold values varied widely within that range, from 0.02 J/cm2 for 285 nm to
 2      0.74 J/cm2 for 315 nm (Oriowo et al., 2001). At wavelengths > 325 nm, the threshold levels
 3      were orders of magnitude larger, with a minimum threshold value of 18.7 J/cm2. A U.S. study of
 4      838 watermen found that UV-B radiation was significantly associated with cortical, but not
 5      nuclear, cataract formation (Taylor et al., 1988). No association was observed between cataracts
 6      and UV-A radiation in this outdoor-working population.
 7
 8      Risk of Ocular Damage from Changes in  Tropospheric Ozone Levels
 9           Cataracts are the most common cause of blindness in the world. McCarty et al. (2000)
10      calculated that ocular UV radiation exposure accounted for 10% of the cortical cataracts in an
11      Australian cohort of 4,744 individuals from both urban and rural areas.  A study by Javitt and
12      Taylor (1994-1995) found that the probability of cataract surgery in the U.S. increased by 3% for
13      each  IE decrease in latitude.  These results suggest that depletion of the stratospheric O3 layer
14      may increase UV radiation-induced cataract formation. After assuming a certain wavelength
15      dependency along with some additional assumptions, every 1% decrease in the stratospheric O3
16      layer was estimated to be associated with a 0.3 to 0.6% increase in cataracts (Longstreth et al.,
17      1995).  Longstreth et al. (1995) noted that this estimate has a high degree of uncertainty due to
18      inadequate information on the action spectrum and dose-response relationships. Quantitative
19      estimates have not been possible for photokeratitis, pterygium, or other UV-related ocular effects
20      due to lack of epidemiologic and experimental data.  As is the case for all of the other UV-
21      related health outcomes, there is no published information on the potential effects of decreased
22      tropospheric O3 concentrations on cataract formation due to any changes in surface-level UV
23      flux resulting from decreases in tropospheric O3.
24
25      10.2.3.5  Ultraviolet Radiation and Immune System Suppression
26           Experimental studies have suggested that exposure to UV radiation may suppress local and
27      systemic immune responses to a variety of antigens (Clydesdale et al., 2001; Garssen and van
28      Loveren, 2001; Selgrade et al., 1997).  In rodent models, these effects have been shown to
29      worsen the course and outcome of some infectious diseases and cancers (Granstein and Matsui,
30      2004; Norval et al., 1999). Granstein and Matsui (2004) stated that exposure to UV-B radiation
31      caused immunosuppression in mice ultimately by releasing cytokines that prevent antigen-

        January 2005                             10-27       DRAFT-DO NOT QUOTE OR CITE

-------
 1      presenting cells from performing their normal functions and causing direct damage to epidermal
 2      Langerhans cells. Noonan et al. (2003) investigated UV skin cancer induction in two strains of
 3      reciprocal Fl hybrid mice and found that genetically determined differences in susceptibility to
 4      UV-induced immunosuppression was a risk factor for skin cancer. At high-UV radiation doses,
 5      mice with greater susceptibility to immune suppression had a larger proportion of skin tumors
 6      compared to those with low susceptibility (Noonan et al., 2003).
 7          While results from animal models are supportive of an association between UV radiation
 8      and local and systemic immunosuppression, evidence for UV-induced immunosuppression in
 9      humans is limited. There is some evidence that UV radiation has indirect involvement in viral
10      oncogenesis through the human papillomavirus (Pfister, 2003). Additional evidence of
11      UV-related immunosuppression comes from an epidemiologic study of 919 patients with rare
12      autoimmune muscle diseases from 15 cities on four continents with variable UV radiation
13      intensity (Okada et al., 2003). Ultraviolet radiation was strongly associated with the prevalence
14      of dermatomyositis, an autoimmune disease distinguished by the presence of photosensitive
15      pathognomonic rashes (Okada et al., 2003). In patients with the human immunodeficiency virus,
16      UV-B radiation lead to activation of the virus in their skin through the release of cytoplasmic
17      nuclear factor kappa B (Breuher-McHam et al., 2001).  In a study by Belgrade et al. (2001),
18      UV-induced immunosuppression was examined in 185  subjects with different skin
19      pigmentations.  To assess immune suppression, dinitrochlorobenzene was applied to irradiated
20      buttock skin 72 hours after irradiation. Differences in sensitivity were not related to skin type
21      based  on the Fitzpatrick classification or minimal erythemal dose (Belgrade et al, 2001).
22      However, erythemal reactivity, assessed by the  steepness of the erythemal dose-response curve,
23      was shown to be significantly associated with UV-induced immunosuppression.  Only subjects
24      with steep  erythemal responses, which included individuals with skin types I through V, showed
25      a dose-response relationship between UV exposure and immune suppression (Belgrade et al,
26      2001).
27          Most action spectrum investigations have concluded that immunosuppression is caused
28      most effectively by the UV-B waveband (Garssen and van Loveren, 2001).  The effects of UV-A
29      on local  and systemic immunosuppression have been unclear and inconsistent. There is some
30      evidence that high doses of UV-A is protective of immunosuppression induced by UV-B
31      exposure (Halliday et al., 2004).  Given the variety of outcomes of immune suppression and

        January 2005                              10-28        DRAFT-DO NOT QUOTE OR CITE

-------
 1      possible mechanisms of effect, little detailed information exists on UV radiation action
 2      spectrums and dose-response relationships.  The available data are insufficient to develop
 3      quantitative risk estimates for UV radiation-induced immunosuppression in humans.
 4
 5      10.2.3.6  Protective Effects of Ultraviolet Radiation - Production of Vitamin D
 6           Most humans depend on sun exposure to satisfy their requirements for vitamin D (Holick,
 7      2004).  UV-B photons are absorbed by 7-dehydrocholesterol in the skin, leading to its
 8      transformation to previtamin D3, which is rapidly converted to vitamin D3. Vitamin D3 is
 9      metabolized in the liver, then in the kidney to its biologically active form.  Vitamin D deficiency
10      is known to cause metabolic bone disease among children and adults, but also may increase the
11      risk of many common chronic diseases, including type I diabetes mellitus and rheumatoid
12      arthritis (Holick, 2004). Vitamin D also is capable of inhibiting the growth of various cancer
13      cells in cell cultures. Thus, vitamin D deficiency may lead to increased incidence of carcinomas
14      in various organs (Studzinski and Moore, 1995).
15           Large geographical gradients in mortality rates for a number of cancers in the U.S. are not
16      explained by dietary or other risk factors; therefore, it has been hypothesized that some
17      carcinomas are due to insufficient UV-B radiation.  For example, published literature indicates
18      that solar UV-B radiation, by increasing production of vitamin D, is associated with reduced risk
19      of cancer of the breast (Freedman et al., 2002; Garland et al., 1990; Gorham et al., 1990; Grant,
20      2002a; John et al., 1999), colon (Freedman et al., 2002), ovary (Freedman et al., 2002; Lefkowitz
21      and Garland, 1994), and prostate (Freedman et al., 2002; Hanchette and Schwartz, 1992), as well
22      as non-Hodgkin lymphoma (Hartge et al., 1996).  Eight other cancers, including bladder,
23      esophageal, kidney, lung, pancreatic, rectal, stomach, and corpus uteri, have been found to
24      exhibit an inverse correlation between mortality rates and UV-B radiation (Grant, 2002b).
25           Using UV-B data from July 1992 and U.S. cancer mortality rates from 1970 to 1994,
26      premature cancer deaths attributable to reduced UV-B exposure were analyzed in an ecologic
27      study (Grant, 2002b). The annual number of premature deaths from various cancers due to
28      lower UV-B exposures was 21,700 for white Americans; 1,400 for black Americans; and 500 for
29      Asian Americans and other minorities. Uncertainty in the estimation of UV-B exposure limits
30      the confidence for the estimates of excess cancer deaths attributable to reduced exposure.
31      No study has assessed the decreased risk of cancer mortality resulting from increased UV

        January 2005                              10-29       DRAFT-DO NOT QUOTE OR CITE

-------
 1      radiation attributable to decreased tropospheric O3 levels, but the change in risk is expected to be
 2      unappreciable.
 3           In establishing guidelines on limits of exposure to UV radiation, the ICNIRP agreed that
 4      some low-level exposure to UV radiation benefits health (ICNIRP, 2004). However, the adverse
 5      health effects necessitated the development of exposure limits for UV radiation.  The ICNIRP
 6      recognized the challenge in establishing exposure limits that would achieve a realistic balance
 7      between beneficial and adverse health effects.
 8
 9      10.2.4  Summary and Conclusions for O3 Effects on UV-B Flux
10           Increased solar radiation, especially UV-B radiation, can be detrimental as well as
11      beneficial to the exposed human population.  Analogously, both positive and negative effects of
12      UV-B radiation would be expected on plant and animal biota, and on man-made materials (e.g.,
13      Van der Leun et al., 1998; U.S. Environmental Protection Agency, 2002). Other environmental
14      factors include the daily nonlinearity of UV-B flux, nonlinear absorption by O3, and absorption
15      and scattering of UV-B radiation by both stratospheric and tropospheric processes, including
16      ground-level environmental pollutants such as PM and O3. Surface flux variability of UV-B is
17      partially dependent upon cloud and tree cover, latitude, time of year, and time of day. Given the
18      magnitude of this natural variability, it would be extremely difficult with current instrumentation
19      and atmospheric models to measure and/or estimate a change in the ground-level UV-B flux and
20      to attribute it to small reductions in troposphericO3 and/or PM.
21           Of equal or even greater importance are changes in human habits of daily activities,
22      recreation, dress, and skin care. Little is known about the impact of variability in these human
23      factors that can modify individual exposure to UV radiation. Most information about potential
24      UV exposure to the human population relies on demographic information. There is some
25      information available on activity patterns (e.g., EPA NHAPS) and protective behaviors (e.g.,
26      Canadian National Survey on Sun Exposure and Protective Behaviors), but the data are
27      inadequate to accurately assess UV radiation exposure in the population. The uncertainties in
28      characterizing UV radiation exposure limit any UV-related health risk assessment.
29           Another difficulty in examining health risks of UV radiation exposure in the population is
30      due to the lack of detailed information regarding the relevant type (e.g., peak or cumulative) and
31      time period (e.g., childhood or lifetime) of exposure,  wavelength dependency of biological

        January 2005                             10-30       DRAFT-DO NOT QUOTE OR CITE

-------
 1      responses, and interindividual variability in UV resistance. For example, exposure to solar UV-
 2      B radiation appears to be the most important environmental risk factor for basal and squamous
 3      cell carcinomas and cutaneous malignant melanoma in fair-skinned individuals. Originally, total
 4      cumulative exposure to UV-B was believed to be the most important environmental factor in
 5      determining risk for these tumors. New information now suggests that only squamous cell
 6      carcinoma risk is related to total accumulated exposure.  For basal cell carcinoma and for
 7      malignant melanoma, new information suggests that increases in risk are linked to early
 8      exposures (before about age  15), particularly those leading to severe sunburns.  There is also
 9      controversy regarding the effect of UV-A on health outcomes. Though most studies have found
10      that UV-B radiation is highly mutagenic, some evidence indicates that UV-A radiation, at high
11      doses, may also be carcinogenic.
12           Susceptibility to the various UV radiation-induced health effects, generally, appears to be
13      related to skin type. Numerous studies have shown that individuals with Fitzpatrick skin types I
14      and II (i.e., fair-skinned phenotypes) are at increased risk for erythema and skin cancer.
15      In addition, individuals with  a reduced ability to repair DNA have increased susceptibility to
16      both nonmelanoma and melanoma skin cancers. For UV-related immunosuppression, however,
17      it appears that erythemal reactivity, rather than skin type, is a marker for susceptibility.
18           There is some evidence that increased UV radiation exposures resulting from depletion of
19      the stratospheric O3 layer will lead to increases in skin cancer and cataracts. However, the
20      numerous assumptions necessary to calculate these quantitative risk estimates lead to high levels
21      of uncertainty. No study has yet examined the benefits of increased UV radiation exposure due
22      to enhanced production of vitamin D. With the  currently available information, the effect of
23      changes in tropospheric O3 on UV-induced health outcomes cannot be critically assessed.
24
25
26      10.3   TROPOSPHERIC OZONE AND CLIMATE CHANGE
27           Water vapor, CO2, O3, N2O, CH4, CFCs, and other polyatomic gases present in the earth's
28      troposphere, trap infrared radiation emitted by the earth's surface, leading to surface warming.
29      This phenomenon is widely known as the "Greenhouse Effect" (Arrhenius, 1896), and the gases
30      involved are known as "greenhouse gases" (GHGs). The term used for the role a particular
31      atmospheric component, or any other component of the greater climate system, plays in altering

        January 2005                              10-31        DRAFT-DO NOT QUOTE OR CITE

-------
 1     the earth's radiative balance is "forcing."  In the past decade, the global atmospheric sciences
 2     and climate communities have made significant progress in determining the specific role O3
 3     plays in forcing climate.
 4           The Intergovernmental Panel on Climate Change (IPCC) was founded in 1988 by the
 5     World Meteorological Society (WMO) and the United Nations Environmental Program (UNEP)
 6     to support the work of the Conference of Parties (COP) to the United Nations Framework
 7     Convention on Climate Change (UNFCCC).  Drawing from the global climate and atmospheric
 8     sciences community for its authors and reviewers, the IPCC produces reports containing
 9     thorough assessments of the available, peer-reviewed science regarding the physical climate
10     system, past and present climate, and evidence of human-induced climate change.  This section
11     will summarize the reviews of the available information on the forcing properties of tropospheric
12     O3 provided by IPCC Third Assessment Report (IPCC, 200la) will be given, and will describe
13     some of the more recent developments on the subject.
14           The projected effects of global climate change will be briefly explained to provide the
15     context within which O3 serves as a regional, and possibly global, anthropogenic pollutant.
16     The concept of climate forcing is also explained, along with the factors that influence the extent
17     of climate forcing by ozone. The section concludes with a summary of the various estimates that
18     have been placed on the globally-averaged forcing due to ozone.
19
20     10.3.1  The Projected Impacts of Global Climate Change
21           The study of the atmospheric processes involved in global climate change and its potential
22     consequences for human health and global ecosystems is an area of active research. The IPCC
23     Third Assessment Report (TAR) is the most thorough evaluation of current scientific
24     understanding of climate change available at this time. In addition to the first and second IPCC
25     assessments in 1990 and 1995, along with other IPCC reports, earlier assessments included those
26     conducted by the United Nations Environment Program (UNEP, 1986), the World
27     Meteorological Organization (WHO, 1988), the U.S. Environmental Protection Agency (1987),
28     and others (e.g., Patz et al., 2000a,b). The reader is referred to these documents for a complete
29     discussion of climate change science. An abbreviated list of the IPCC conclusions, to date, and a
30     short discussion of the potential impacts of climate change on human health and welfare is
       January 2005                             10-32      DRAFT-DO NOT QUOTE OR CITE

-------
 1      provided here to serve as the context for the discussion of the role of the increasing tropospheric
 2      O3 concentration in climate change.
 3           According to various historic and modern measurement records, atmospheric GHG
 4      concentrations have increased dramatically in the past century, and have been attributed to
 5      human activities.  The IPCC TAR describes the scientific theory and evidence tying increases
 6      in GHGs to human  activities (IPCC 2001 a).
 7           An increasing body of geophysical observations shows that the earth is warming and that
 8      other climate changes are underway.  These observations include the global surface temperature
 9      record assembled since the year 1860, the satellite temperature record begun in 1979, recorded
10      changes in snow and ice cover since the 1950s, sea level measurements taken throughout the
11      20th century, and sea surface temperature observations recorded since the 1950s.  Other
12      evidence includes a marked increase over the past 100 years in the frequency, intensity, and
13      persistence of the zonal atmospheric circulation shifts known as the El Nino-Southern
14      Oscillation (ENSO). ENSO events occur when the tropical Pacific Ocean has accumulated a
15      large, localized mass of warm water that interrupts cold surface currents along South America,
16      altering precipitation and temperature patterns in the tropics, subtropics and the midlatitudes.
17           IPCC (1998, 200la) reports also describe the results of general circulation model (GCM)
18      studies that predict  that human activities will alter the climate system in a manner that will likely
19      lead to marked global and regional changes in temperature, precipitation and other climate
20      properties. These changes are expected to increase the global mean sea level, as well as increase
21      the number of extreme weather events including floods and droughts, increased wind speeds and
22      precipitation intensity of tropical cyclones, and changes in soil moisture.  These predicted
23      changes can be expected to directly impact human health, ecosystems,  and global  economic
24      sectors, e.g., hydrology and water resources, food and fiber production (IPCC, 1998, 200Ib).
25      Table 10-1 summarizes these projected impacts.
26           Wide variations in the course and net impacts of climate change in different geographic
27      areas are expected.  In general, the projected changes will result in additional stresses on natural
28      ecosystems and human societal systems already impacted by increasing resource demands,
29      unsustainable resource management practices, and pollution. Some of the predicted changes
30      include alterations in ecological balances; in the availability of adequate food, water, clean air;
        January 2005                              10-33       DRAFT-DO NOT QUOTE OR CITE

-------
      Table 10-1. Examples of Impacts Resulting From Projected Changes in Extreme
                                         Climate Events
 Projected changes during the 21st Century in
 Extreme Climate Phenomena and their
 Likelihood3

 Simple Extremes

 Higher maximum temperatures; more hot days
 and heat waves'1 over nearly all land areas (very
 likely^)
 Higher (increasing) minimum temperatures;
 fewer cold days, frost days, and cold waves'1
 over nearly all land areas (very likely3)
 More intense precipitation events (very likely*
 over many years)
 Complex Extremes

 Increased summer drying over most midlatitude
 continental interiors and associated risk of
 drought (likely*)
 Increase in tropical cyclone peak wind
 intensities, mean and peak precipitation
 intensities (likely* over some areas)6
 Intensified droughts and floods associated with
 El Nino events in many different regions
 (likely*) (see also under droughts and intense
 precipitation events)

 Increased Asian summer monsoon precipitation
 variability (likely*)
Representative Examples of Projected Impacts*
(all high confidence of occurrence in some areas0)
  Increased incidence of death and serious illness in
  older age groups and urban poor
  Increased heat stress in livestock and wildlife
  Shift in tourist destinations
  Increased risk of damage to a number of crops
  Increased electric cooling demand and reduced energy
  supply reliability

  Decreased cold-related human morbidity and mortality
  Decreased risk of damage to a number of crops, and
  increased risk to others
  Extended range and activity of some pest and disease
  vectors
  Reduced heating energy demand

  Increased flood, landslide, avalanche, and mudslide
  damage
  Increased soil erosion
  Increased flood runoff could increase recharge of
  some floodplain aquifers
  Increased pressure on government and private flood
  insurance systems and disaster relief
  Decreased crop yields
  Increased damage to building foundations caused by
  ground shrinkage
  Decreased water resource quantity and quality
  Increased risk of forest fire

  Increased risk to human life, risk of infections, disease
  epidemics, and many other risks
  Increased coastal erosion and damage to coastal
  buildings and infrastructure
  Increased damage to coastal ecosystems such as coral
  reefs and mangroves

  Decreased agricultural and rangeland productivity in
  drought- and flood-prone regions
  Decreased hydro-power potential in drought-prone
  regions

  Increased flood and drought magnitude and damages
  in temperate and tropical Asia
January 2005
10-34
DRAFT-DO NOT QUOTE OR CITE

-------
             Table 10-1 (cont'd). Examples of Impacts Resulting From Projected Changes in
        	Extreme Climate Events	
         Projected changes during the 21st Century in     Representative Examples of Projected Impacts*
         Extreme Climate Phenomena and their          (all high confidence of occurrence in some areas0)
         Likelihood3
         Increased intensity of midlatitude storms        • Increased risks to human life and health
         (little agreement between current models)d       • Increased property and infrastructure losses
                                                  • Increased damage to coastal ecosystems

         "Likelihood refers to judgmental estimates of confidence used by TAR WGI: very likely (90-99% chance); likely
         (66-90% chance). Unless otherwise stated, information on climate phenomena is taken from the Summary for
         Policymakers, TAR WGI. TAR WGI = Third Assessment Report of Working Group 1 (IPCC,  2001a).
         bThese impacts can be lessened by appropriate response measures.
         "High confidence refers to probabilities between 67 and 95%.
         Information from TAR WGI, Technical Summary.
         eChanges in regional distribution of tropical cyclones are possible but have not been established.
         Source:  IPCC(2001b).
 1      and in human health and safety. Poorer nations can be expected to suffer the most, given their
 2      limited adaptive capabilities.
 3           Although many regions are predicted to experience severe, possibly irreversible, adverse
 4      effects due to climate change, beneficial changes may also take place. For example, certain
 5      regions may benefit from warmer temperatures or increased CO2 fertilization, e.g., U.S. West
 6      Coast coniferous forests, some Western rangelands.  Specific benefits may include reduced
 7      energy costs for heating, reduced road salting and snow-clearance costs, longer open-water
 8      seasons in northern channels and ports, and improved agricultural opportunities in the northern
 9      latitudes, as well as the Western interior and coast.  For further details about the projected effects
10      of climate change on a U.S.-regional scale, the reader is also referred to several regionally-
11      focused reports (MARAT, 2000; Yarnal et al., 2000; NERAG, 2001; GLRAG, 2000), as well as
12      a report on potential impacts upon human health due to climate change (Bernard et al.,
13      2001).The IPCC report, "The Regional Impacts of Climate Change," (IPCC, 1998) describes the
14      projected effects of human-induced climate change on the different regions of the globe,
15      including Africa, the Arctic and Antarctic, the Middle East and arid Asia, Australasia, Europe,
16      Latin America, North America, the small island nations, temperate Asia, and tropical Asia.
17      While climate models are successful in simulating present annual mean global climate and the

        January 2005                              10-3 5       DRAFT-DO NOT QUOTE OR CITE

-------
 1      seasonal cycles on continental scales, they have been less successful on regional scales. Clouds
 2      and humidity, essential factors in defining local and regional (sub-grid scale) climate, are
 3      significantly uncertain (IPCC, 200la).  Due to modeling uncertainties, both in reproducing
 4      regional climate and in predicting the future economic activity that will govern future GHG
 5      emissions, the projected impacts discussed above are also uncertain.
 6           Findings from the United States Global Change Research Program (USGCRP) (NAST,
 7      2000) report and related reports illustrate the considerable uncertainties and difficulties in
 8      projecting likely climate change impacts on regional or local scales. The USGCRP findings also
 9      reflect the mixed nature of projected potential climate change impacts (i.e., combinations of
10      mostly deleterious, but other possibly beneficial, effects) for U.S. regions and their variation
11      across different regions. Difficulties in projecting region-specific climate change impacts are
12      complicated by the need to evaluate the potential effects of regional- or local-scale changes in
13      key air pollutants not only on global-scale temperature trends, but also regional- or local-scale
14      temperature and precipitation patterns.
15
16      10.3.2  Solar Energy Transformation and the Components of the Earth's
17              Climate System
18           Mass, in any form, has the capacity to interact with solar and terrestrial radiation, but the
19      manner in which mass interacts with radiation is governed by its particular physical form and/or
20      molecular properties. The most interesting example of radiative properties as a function of
21      physical form is water. In its gas phase, water is the most important GHG present in the climate
22      system, due to its ability to absorb long-wave terrestrial radiation.  Conversely, in its frozen form
23      as snow or sea ice, the most important role for water in the climate system is scattering
24      ultraviolet and visible solar radiation back to space, i.e., decreasing the earth's net solar radiation
25      receipts by increasing the earth's reflective properties (albedo). In its  liquid aerosol form as
26      clouds, water scatters radiation.  In its bulk liquid form as ocean water, it absorbs terrestrial
27      radiation, and represents the earth's most important reservoir of heat energy due to its mass.
28      Molecular properties become important in comparing the  radiative absorptive capacity of
29      different gases. In the case of gases, the atomic composition and molecular structure determine
30      the wavelengths of light that a gas molecule can absorb.  Ozone and O2, provide one example of
31      the importance of molecular structure in determining absorption capacity.  These molecules are

        January 2005                              10-36      DRAFT-DO NOT QUOTE OR CITE

-------
 1      composed solely of atomic oxygen atoms, but their bond structures are distinct.  Ozone, because
 2      of its three-atom, bent molecular structure, has the capacity to absorb terrestrial (infrared)
 3      wavelengths, making it a GHG. Tropospheric O3, like stratospheric O3, also has the capacity to
 4      absorb ultraviolet radiation of 320 nm and shorter, increasing the energy-absorbing capacity of
 5      the troposphere. Molecular oxygen, O2, due to its structure, is limited to absorbing very short-
 6      wave ultraviolet light - and does so at altitudes too high to influence the climate system
 7      significantly.
 8           Each component of the climate system plays a role in absorbing, transforming, storing,
 9      dispersing, or scattering solar radiation.  Weather is a consequence of the transformation and
10      dispersion of terrestrial radiation.  The term, "weather," refers to the condition of the earth's
11      atmosphere at a given time and place. It is defined by the air temperature, air pressure, humidity,
12      clouds, precipitation, visibility, and wind speed. The "climate" for a given place on the earth's
13      surface is a long-term average of these elements of weather, accounting for daily and seasonal
14      weather events. The frequency of extreme weather events is used to distinguish among climates
15      that have similar averages (Ahrens, 1994).
16           The earth's capacity for retaining heat is increased by the transformation of its oxygen, for
17      example, into O3 by  way of air pollution chemistry. Climate components, including GHGs, land,
18      oceans, sea ice, land ice and snow, atmospheric particles, vegetation, clouds, etc., all contribute
19      to the earth's heat capacity with respect to solar energy.  Changes in the properties (or mass) of
20      these components will "force" the climate system in one direction or the other, i.e., warmer
21      versus cooler.  Climate forcing is further described, below.
22
23      10.3.3   The Composition of the Atmosphere and the Earth's Radiative
24               Equilibrium
25           The Greenhouse Effect is the term given to the decreased rate of reemission of absorbed
26      solar energy due to the heat-retaining properties of the Earth's atmosphere.  According to simple
27      radiative transfer theory, at thermal equilibrium, the earth's temperature  should be near -15 °C.
28      This is the temperature of a theoretical "black body" that is receiving and then reemitting
29      342.5 WnT2, i.e., the globally-averaged amount of full-spectrum solar energy  absorbed, and then
30      reemitted by the earth as infrared terrestrial radiation, per square meter.  In fact,  satellite
31      observations well above the atmosphere indicate that the earth's average planetary temperature

        January 2005                              10-37       DRAFT-DO NOT QUOTE OR CITE

-------
 1      is remarkably close to its theoretical black body value at -18 °C, a temperature at which liquid
 2      water ordinarily does not exist.
 3           At its surface, however, the earth's average temperature is +15 °C.  The +33 °C
 4      temperature differential between the Earth's planetary and surface temperatures is due to the
 5      presence of infrared radiation-absorbing components in the atmosphere, such as water vapor,
 6      CO2, CH4, several other trace gases, and some types of particles and clouds.
 7           The atmosphere, when cloud free, is largely transparent in the solar wavelength range.
 8      A small fraction of this radiation is absorbed and reemitted as black body radiation by dark
 9      atmospheric particles (IPCC, 200la).  The majority of clouds and particles offset, in part, the
10      greenhouse effect by increasing the earth's albedo and, therefore, decreasing the overall  amount
11      of solar radiation absorbed by the earth system.
12           Ozone, SO2 and NO2 also absorb ultraviolet and near ultraviolet wavelengths, in addition to
13      infrared wavelengths. Once absorbed by a gas molecule, the energy introduced by a photon may
14      induce  a photochemical reaction with the residual energy thermally exciting (heating) the
15      products of the reaction.  Alternatively, the energy introduced into the molecule by the photon
16      may be dispersed amongst neighboring molecules via intermolecular collisions, or reemitted in
17      part as  a lower energy (IR) photon.
18           Radiation from the sun or the earth's surface that is absorbed by gases and particles is
19      reemitted isotropically, i.e., it is equally likely to be emitted in all directions. Therefore, to a
20      first approximation, half of the radiation trapped by the earth's atmosphere is reflected back to
21      its surface. A portion of this radiation is transformed into the heat energy that drives the
22      atmospheric processes that form the basis of weather and climate. Radiation which is not
23      absorbed by gases and aerosols reaches the earth's surface where it is scattered (reflected) or
24      absorbed, depending on the surface albedo.
25           Successful modeling of the earth's climate and, therefore, the assessment of the extent of
26      human-induced climate change and development of appropriate policy depends on the quality of
27      available information on the relative efficiencies, amounts, and spatial and temporal distributions
28      of the various radiatively active components of the atmosphere  that absorb and/or reflect solar
29      and terrestrial radiation, along with all the other non-atmospheric components of the earth
30      system.
31

        January 2005                              10-3 8       DRAFT-DO NOT QUOTE OR CITE

-------
 1      10.3.3.1  Forcing of the Earth's Radiative Balance
 2           As mentioned earlier, the commonly used measure of the relative influence of a given
 3      component of the climate system on the earth's radiative balance is its radiative forcing (IPCC,
 4      2001a; Houghton et al.,  1990). Radiative forcing, in WnT2, is a quantity that was developed by
 5      the climate modeling community as a first order-only means of estimating relative effects of
 6      individual anthropogenic and natural processes on the energy balance within the climate system.
 7      Discussions within the climate community are underway regarding a metric to replace forcing as
 8      the standard measure of climate impact - one that will account for more of the factors that
 9      determine the effectiveness of a given atmospheric component in this capacity. However,
10      forcing remains the current standard (National Research Council, 2005).
11           When the effect of a particular component of the climate system is to reduce the amount of
12      solar energy absorbed, usually by increasing the earth's albedo, this component is said to provide
13      a "negative" forcing, measured in WnT2. The convention assigns a positive value to the forcing
14      induced by climate system components that enhance the greenhouse effect, or otherwise act to
15      increase the heat absorbing capacity of the earth system. Purely reflective atmospheric aerosol,
16      clouds, white rooftops,  snow-covered land surfaces, and dense sea ice provide a negative
17      forcing, while highly absorbing dark-colored atmospheric aerosols, GHGs, and increases, due to
18      the melting of sea ice sheets, in dark ocean surfaces positively force the climate system.
19           Global and regional climate are  roughly defined by the balance between the large number
20      of positive and  negative forcings induced by the many different components of the earth system.
21      However, the earth system responds to these forcings in a complex way, due to feedback
22      mechanisms that are theorized but very difficult to resolve at the quantitative level.
23           A simple  example would be the positive feedback associated with melting sea ice. As sea
24      ice melts with increasing surface temperatures, the dark ocean surface is revealed which is more
25      efficient at absorbing infrared radiation, further increasing the rate of warming. The role of
26      feedbacks in determining the sensitivity of climate to changes in radiative forcing is described in
27      detail in the IPCC TAR (IPCC, 200la). In the absence of complete, quantitative, information
28      about climate feedbacks the use of radiative forcing values for the  many components of the
29      climate system  remains the primary method for estimating their relative importance in climate
30      change.
31

        January 2005                              10-39       DRAFT-DO NOT QUOTE OR CITE

-------
 1
 2
 3
 4
 5
 6
 7
 8
 9
10
11
12
     The IPCC has reported estimated values for forcing by individual radiatively active gases,
and particle-phase components of the atmosphere, derived primarily through expert judgment
incorporating the results of peer-reviewed modeling studies. The forcing estimates, shown in
Figure 10-5, are global averages attributed to known greenhouse gases, including O3; particles;
anthropogenic cirrus clouds; land-use change; and natural solar flux variations. Uncertainty
ranges are assigned to reflect the range of modeled values reported in those studies. The current
estimate of forcing due to long-lived, well-mixed, greenhouse gases accumulated in the
atmosphere from the preindustrial era (ca., 1750) through the year 2000 is + 2.4 WnT2 ± 10%
(IPCC, 2001a).  An indication of the level of confidence in each of these estimates is given along
the bottom of this figure, again reflecting the expert judgment of the IPCC.
               CD
               *j
               V
                    2-
               re  .g
               =  E
               
-------
 1           The IPCC reported a global average forcing value of 0.35 ± 0.15 Wrrf2 for tropospheric O3,
 2      based on model calculations constrained by climatological observations. The considerations and
 3      studies used to estimate this value will be outlined below.
 4
 5      10.3.4   Factors Affecting the Magnitude of Climate Forcing by Ozone
 6           Tropospheric O3 is estimated to have provided the third largest increase in direct forcing
 7      since pre-industrial times.  It may also play a role in indirect forcing through its participation in
 8      the oxidative removal of other radiatively active trace gases, such as CH4 and the HCFCs. Given
 9      its relatively short atmospheric lifetime, the distribution of tropospheric O3 is highly variable in
10      geographic extent and time. The direct and indirect forcing it imposes on the climate system,
11      therefore, depends upon its geospatial and temporal distribution, but also depends upon the
12      albedo of the underlying surface. These several variables introduce substantial uncertainty into
13      tropospheric ozone forcing estimates.
14
15      10.3.4.1   Global versus Regional Atmospheric Ozone Concentrations
16           Ozone reacts photochemically at time-scales generally shorter than those for large-scale
17      mixing processes in the atmosphere.  Concentrated O3 plumes evolve downwind of strong
18      sources of its precursor pollutants:  reactive nitrogen, CO, and non-methane hydrocarbons
19      (NMHCs). The most important of these sources are midlatitude industrialized areas and tropical
20      biomass burning. When viewed from above the atmosphere by satellite-borne spectrometers, O3
21      enhancements appear as relatively localized air masses or regional-scale plumes, usually
22      originating from industrialized areas or areas in which active biomass burning is underway. The
23      IPCC (200la) describes the efforts of several research teams who have analyzed data supplied by
24      the satellite-borne Total Ozone Mapping Spectrometer (TOMS) and other remote-sensing
25      instruments to map the global distribution of tropospheric O3 and to attempt to identify processes
26      that influence the global tropospheric O3 budget (IPCC, 200la). More recently, coincident
27      observations of total O3 by TOMS and the Solar Backscattered  UV (SBUV) instrument were
28      used by Fishman et al.  (2003) to construct well-resolved spatial and temporal maps of the
29      regional distribution of tropospheric O3.  Their results were consistent with those reported by
30      others, but with higher regional-scale resolution. They reported large O3 enhancements in the
31      southern tropics in austral Spring, and in the northern temperate latitudes in the summer.  The

        January 2005                              10-41      DRAFT-DO NOT QUOTE OR CITE

-------
 1      regional nature of high O3 concentrations was clearly visible in northeastern India, the eastern
 2      United States, eastern China, and west and southern Africa, each coincident with high population
 3      densities. Fishman et al. (2003) noted, as have the other groups cited above, significant
 4      interannual variability in the concentrations observed over these regions. In situ measurements
 5      of tropospheric O3 concentrations range from 10 ppb over remote oceans, to 100 ppb in both the
 6      upper troposphere and in plumes downwind from polluted metropolitan regions (IPPC, 200la).
 7      Ground-level concentrations in urban areas are often higher than 100 ppb. In the southern
 8      hemisphere, one of the largest  sources of O3 precursors is biomass burning. Biomass burning
 9      elevates O3 on large spatial scales, particularly in the tropical Atlantic west of the coast of Africa
10      and in Indonesia.
11           Current estimates place the global burden of tropospheric O3 at a highly uncertain
12      370 Tg, equivalent to an average column density of 34 Dobson Units (1 DU = 2.687 x 1016
13      molecules/cm"2) or a mean concentration of 50 ppb (IPCC, 200la).  Accounting for differences
14      in levels of industrialization between the hemispheres, the average column burden in the
15      Northern Hemisphere is estimated to be 36 DU, with the Southern Hemisphere estimated to
16      average 32 DU. Due to its rapid photochemistry, individual surface measurements of
17      tropospheric O3 cannot capture large scale concentrations, nor will it represent the concentration
18      at higher altitudes. Dense surface and vertical measurements (ozonesondes) would be required
19      to supplement available output from remote sensing instruments to provide the complete set of
20      observations needed to derive a credible global  O3 budget. Such a measurement program
21      appears, at present, to be impractical.
22           Little historical data exists that might be used to estimate the global ozone burden prior to
23      industrialization.  Although a few late  19th century measurements suggest that O3 has more than
24      doubled in Europe during the 20th century. The insufficient data record on pre-industrial
25      tropospheric O3 distributions introduces a major uncertainty in the estimation of the ozone-
26      induced forcing (IPCC, 200la).
27
28      10.3.4.2  Global Versus  Regional Atmospheric Ozone Trends
29           For the northern hemisphere, weekly  continuous data since 1970  are available from only
30      nine stations in the latitude range 36°N to 59°N (IPCC,  2001a).  Available tropospheric O3
31      measurements do not reveal a clear trend in concentration, while trends in the stratosphere are

        January 2005                              10-42       DRAFT-DO NOT QUOTE OR CITE

-------
 1
 2
 3
 4
 5
 6
 7
 8
 9
10
more readily identified. Different trends are seen at different locations for different periods,
consistent with regional changes in pollutant emission, especially NOX. Logan et al. (1999)
analyzed the composite record of mid-tropospheric ozone abundance from the nine station
network. A plot of data is shown in Figure 10-6.  While no clear trend appeared for 1980
through 1996, the second half of this record (about 57 ppb) is clearly greater than the first half
(about 53 ppb).  The trend may be consistent with changes in regional NOx emission rates due to
pollution reduction efforts in developed countries and increasing emissions in rapidly growing
economies in Asia.
                         70
                         65
                         60
                      2"
                      Q.
                      S 55
                      O
                         50
                         45

                         40
                               400 to 630 hPa
                               36°N to 59°N
                            1970
                                     1975
                                              1980
                                                       1985      1990     1995
       Figure 10-6.  Mid-tropospheric O3 abundance (ppb) in northern midlatitudes
                     (36 °N-59 °N) for the years 1970 to 1996. Observations between 630 and
                     400 hPa are averaged from nine ozone sonde stations (four in North
                     America, three in Europe, two in Japan), following the data analysis of
                     Logan et al. (1999). Values are derived from the residuals of the trend fit
                     with the trend added back to allow for discontinuities in the instruments.
                     Monthly data (points) are shown with a smoothed 12-month-running
                     mean  (line).
       Source: IPCC(2001).
 1          It must be noted that the measurements shown in Figure 10-6 are for surface
 2     concentrations, only. Many fewer locations have measured changes in the concentrations of O3
 3     as a function of altitude. Fewer still are locations that have collected and maintained data
       January 2005
                                         10-43
DRAFT-DO NOT QUOTE OR CITE

-------
 1     records prior to 1970.  The absence of historical data on the vertical distribution of O3 adds to the
 2     difficulty in estimating historical atmospheric burdens and trends in O3-related climate forcing.
 3          The IPCC (200la) surveyed the results of published chemistry transport model (CTM)
 4     modelling studies, listed in Table 10-2, that estimated the global average increase since the
 5     pre-industrial era in total column O3.  Model estimates ranged from +7 to +12 DU.  On the basis
 6     of these estimates, available measurements and other analyses, the IPCC estimated that total
 7     column O3 has increased by 9 DU, with a 67% confidence  range of+6 to +13 DU. In some of
 8     the modelling studies,  emissions scenarios were used that predict a further increase in column
 9     O3 due to growing emissions of O3 precursors.
10
11
            Table 10-2.  CTM Studies Assessed by the IPCC for its Estimate of the Change in
                         Global, Total Column  O3  Since the Pre-industrial Era
Estimated Change in Column
Ozone in DU
7.9
8.9
8.4
9.5
12
7.2
8.7
9.6
8
Model Used
GFDL
MOZART-1
NCAR/2D
GFDL-scaled
Harvard/GISS
ECHAM4,
UKMO,
UIO
MOGUNTIA
Authors (Publication Date)
Haywoodetal. (1998)
Hauglustaine et al. (1998)
Kiehletal. (1999)
Levyetal. (1997)
Mickley et al. (1999)
Roelofsetal. (1997)
Stevenson et al. (2000)
Berntsen et al. (1999)
VanDorland et al. (1997)
        Source: IPCC (2001).
 1          Chapters 2 and 3 of this document provide an expanded discussion of the issues associated
 2     with determining the global tropospheric O3 background.
 3
       January 2005                             10-44       DRAFT-DO NOT QUOTE OR CITE

-------
 1      10.3.4.3  The Sensitivity of Ozone-related Forcing Surface to Albedo
 2           The characteristics of the surface underlying an O3 enhancement play a large role in the O3
 3      forcing effect. Highly reflective, warm surfaces, such as light-colored deserts, scatter solar
 4      ultraviolet (UV) radiation; absorb, and then emit infrared terrestrial radiation. Both forms of
 5      radiation can be trapped, transformed, and/or reemitted back to the surface by tropospheric O3.
 6      Highly reflective, cold, surfaces will scatter more radiation while emitting less terrestrial infrared
 7      radiation.  Dark, warm, surfaces such as tropical ocean hot spots predominantly emit terrestrial
 8      radiation.  Ozone will trap heat at differing efficiencies as a consequence of the amount and type
 9      of radiation reflected or emitted from the surface underneath it.  Studies by two groups,
10      Hauglustaine et al. (1998) and Mickley et al. (1999), have shown that industrial pollution that
11      has been transported to the Arctic induces a high, regional O3-related forcing due to the highly
12      reflective underlying ice and snow surface.
13
14      10.3.5  Estimated Forcing by Tropospheric Ozone
15      10.3.5.1  Direct Climate Forcing Due to Ozone
16           The inhomogeneous distribution of O3 within the troposphere coupled with the large
17      uncertainty in the global O3 budget significantly complicates the matter of estimating the global
18      average direct forcing due to O3. The  IPCC TAR (200la) lists the results of several modeling
19      studies that estimated the annual change in the relative forcing by O3 since pre-industrial times.
20      It was noted that the differences amongst the estimates were most likely due to differences in
21      predicted O3 chemistry, including the emissions inventories used, and the chemical process and
22      transport mechanisms incorporated into the models, rather than by factors relating to radiative
23      transfer. The IPCC intercomparison of the models and their results indicated that the
24      uncertainties in estimated forcings due to O3 have reduced since the  IPPC Second Assessment
25      Report (1996)
26           The O3-related forcings estimated by studies considered by the IPCC (200la) are listed in
27      Table 10-3. Ten of the listed estimates are based on global chemistry/transport models
28      calculations. One study was constrained by a climatology derived from observations. Given the
29      differences in calculated total column  O3 amongst the models, a normalized forcing (WnT2 per
30      Dobson Unit of tropospheric O3 change) is listed in addition to the absolute forcing (WnT2)
        January 2005                              10-45       DRAFT-DO NOT QUOTE OR CITE

-------
          Table 10-3. Tropospheric O3 Change (AO3) in Dobson Units (DU) Since Pre-industrial
              Times, and the Accompanying net (SW plus LW) Radiative Forcings (Wm~2),
              After Accounting for Stratospheric Temperature Adjustment (using the Fixed
           Dynamical Heating method). Estimates are Taken From the Published Literature.
        Normalized Forcings (norm.) Refer to Radiative Forcing per O3 Change (Wm~2 per DU)
Estimated Global Average Forcing Due to TronosDheric Ozone
Clear sky conditions
Reference
Berntsen et al. (1997) - [Reading model]
Stevenson etal. (1998)
Berntsen et al. (1997) - [Oslo model]
Haywoodetal. (1998a)
Kiehl etal. (1999)
Berntsen et al. (2000)
Brasseur etal. (1998)
van Dorland et al. (1997)
Roelofs etal. (1997)
Lelieveld and Dentener (2000)
Hauglustaine et al. (1998)
Mean
iO3
7.600
8.700
7.600
7.900
8.400
9.600


8.070
7.200
—
8.940
8.224
Net
0.310
0.391
0.390
0.380
0.379
0.428


0.443
0.397
—
0.511
0.403
Net (norm.)
0.041
0.045
0.051
0.048
0.045
0.045


0.055
0.055
—
0.057
0.049
Total sky conditions
Net
0.280
0.289
0.310
0.310
0.320
0.342
0.370
0.380
0.404
0.420
0.426
0.343
Net (norm.)
0.037
0.033
0.041
0.039
0.038
0.036


0.047
0.056
—
0.048
0.042
        Source: IPCC(2001).
1     estimated by each model. Both clear sky (cloud-free) and total sky (including clouds) forcing
2     estimates are listed.
3          The largest O3-related forcings coincide with strongest sources of tropospheric ozone, the
4     models predict that occur in the northern midlatitude regions (40° to 50° N), reaching as much as
5     1 WnT2 in the summer and in the tropics, related to biomass burning. In general, the estimates
6     are comparable in magnitude and show similarity in geographic distribution.  For total sky
7     conditions, the range in globally and annual averaged tropospheric O3 forcing from all of these
8     models is from 0.28 to 0.43 WnT2, while the normalized forcing is 0.033 to 0.056 WnT2 per DU.
9     As expected, they are opposite in sign to the forcing estimated for sulphate aerosols, which
      January 2005
10-46
DRAFT-DO NOT QUOTE OR CITE

-------
 1      scatter radiation.  The range in normalized forcings emphasizes the differences in assumptions
 2      used by the different models.  The tropospheric O3 forcing constrained by the observational
 3      climatology is 0.32 WnT2 for globally averaged, total sky conditions.  As shown in Figure 10-1,
 4      the IPCC (2001) concluded that 0.35 ± 0.15 WnT2represents the most likely value for annually
 5      and globally-averaged forcing by tropospheric O3.
 6
 7      10.3.5.2 Indirect Forcing Due to Ozone
 8           Ozone has an indirect climate forcing effect due to its role in the oxidative removal of
 9      other reactive GHGs, including CH4, hydrofluorocarbons (HFCs) and other reactive
10      non-methane hydrocarbons (NMHCs). The primary actor in this effect is a second generation
11      product of the photolysis of O3, the hydroxyl radical (OH). Hydroxyl radical is produced by
12      way of a pair of reactions that start with the photodissociation of O3 by solar UV.
13
14
                                       O3 + hv ->• O(1D) + O2                               (10-5)
16
17                                   O(1D) + H2O -> OH + OH                             (10-6)
18
19      Reactions with OH are the primary removal mechanism for CH4 and NMHCs as well as the
20      pollutants NOX and CO.  Methane and CO have especially high abundances in the global
21      atmosphere.  OH is estimated to react with these two gases within one second of its formation.
22      In addition to CH4, NOX, CO and the NMHCs, OH concentrations are controlled by local
23      concentrations  of H2O (humidity) and the intensity of solar UV. Different atmospheric
24      concentrations  of the required precursors suggest that pre-industrial OH concentrations are likely
25      to have been different from present-day concentrations, but there is no consensus on  the
26      magnitude of this difference. Observations of global atmospheric concentrations of chloroform
27      (CH3CC13), a well-mixed tropospheric species that reacts with OH, have been used to estimate
28      OH abundances.  Independent studies have shown overlapping trends for the period  1978 to
29      1994, but none outside the given uncertainty ranges (0.5 ± 0.6%/yr) (Prinn et al., 1995; Krol
30      et al., 1998).  The IPCC (2001) reported a range of+5% to -20% for predicted changes in global
31      OH abundances.
        January 2005                              10-47       DRAFT-DO NOT QUOTE OR CITE

-------
 1           Given the difficulty in estimating global OH abundances in the past, present and future,
 2      estimates of indirect forcing due to O3 have been difficult to obtain and are highly uncertain.
 3      Attempts have been made to account for OH feedbacks in estimating the global warming
 4      potential of CH4.
 5
 6      10.3.5.3  Predictions for Future Climate Forcing by Anthropogenic Ozone
 7           The rate of increase in surface O3 in Europe and North America, since 1980, appears to be
 8      slowing, likely due to control measures intended to improve urban air quality.  Not surprisingly,
 9      CTM modeling attempts to predict future precursor emissions and resulting O3 abundances
10      indicate that the largest future O3-related forcings will be related to population growth and
11      economic development in Asia (van Borland et al., 1997; Brasseur et al., 1998). The results of
12      these modeling studies suggested that a higher globally averaged total radiative forcing due to O3
13      from pre-industrial times to 2050 of 0.66 WnT2 and 0.63 WnT2. Chalita et al. (1996) predicted a
14      globally averaged  radiative forcing from pre-industrial times to 2050 of 0.43 WnT2. Stevenson
15      et al. (1998) predicted an O3-related forcing of 0.48 WnT2 in 2100. All of these predictions must
16      be viewed with much caution given the considerable  uncertainties associated with such
17      estimates.
18
19      10.3.6  Conclusion
20           The general consensus within the atmospheric sciences community, as represented by the
21      United Nations Intergovernmental Panel on Climate Change (IPCC), is that human activities
22      have a discernable effect on the earth's climate. However, quantifying the extent of human-
23      induced forcing on climate  requires detailed information about human-induced change on the
24      components of the earth system that govern climate.  Troposphere ozone is a well-known GHG,
25      but information regarding its historical trends in concentration, its current  and future
26      atmospheric burden, and other critical details needed for estimating its direct and indirect
27      forcing effects on the climate system are highly uncertain.
28           The IPCC has estimated that the globally averaged forcing due to O3 is approximately
29      0.35 ± 0.15  WnT2. The role of O3 in climate is likely to be much more pronounced adjacent to
30      the sources  of its chemical precursors, consistent with satellite observations of high regional
31      scale column densities near large urban areas and large-scale biomass burning activity.

        January 2005                              10-48       DRAFT-DO NOT QUOTE OR CITE

-------
1     Modeling studies evaluated by the IPCC have estimated that regional scale forcing due to O3 can
2     approach 1 WnT2, or two-thirds the forcing estimated for CO2. However, more reliable
3     estimates of the overall importance of forcing due to tropospheric O3 await further advances in
4     monitoring and chemical transport modeling.
5
      January 2005                             10-49       DRAFT-DO NOT QUOTE OR CITE

-------
  1       REFERENCES

  2       Ahrens, D. C. (1994) Meteorology today: an introduction to weather, climate and the environment. Minneapolis,
  3             MN: West Publishing Co., p. 17.
  4       Airey, D. K.; Wong, J. C.; Fleming, R. A.; Meldrum, L. R. (1997) An estimate of the UV-B exposure for outdoor
  5             workers during a south-east Queensland summer. Health Phys. 72: 544-549.
  6       Ancellet, G.; Beekmann, M. (1997) Evidence for changes in the ozone concentrations in the free troposphere over
  7             southern france from 1976 to 1995. Atmos. Environ. 31: 2835-2851.
  8       Autier, P.; Dore, J. F.; Negrier, S.; Lienard, D.; Panizzon, R.; Lejeune, F. J.; Guggisberg, D.; Eggermont, A. M.
  9             (1999) Sunscreen use and duration of sun exposure: a double-blind, randomized trial. J. Natl. Cancer Inst.
10             91: 1304-1309.
11       Autier, P.; Dore, J.-F.; Reis, A. C.; Grivegnee, A.; Ollivaud, L.; Truchetet, F.; Chamoun, E.; Rotmensz, N.;
12             Severi, G.;  Cesarini, J. P.; EORTC Melanoma Co-operative Group. (2000) Sunscreen use and intentional
13             exposure to ultraviolet A and B radiation: a double blind randomized trial using personal dosimeters. Br. J.
14             Cancer 83:  1243-1248.
15       Ayala, M. N.; Michael, R.; Soderberg, P. G. (2000) Influence of exposure time forUV radiation-induced cataract.
16             Invest. Ophthalmol. Visual Sci. 41: 3539-3543.
17       Barnard, W. F.; Saxena, V. K.; Wenny, B. N.; DeLuisi, J. J. (2003) Daily surface UV exposure and its relationship
18             to surface pollutant measurements. J. Air Waste Manage. Assoc. 53: 237-245.
19       Bellows, C. F.; Belafsky, P.; Fortgang, I. S.; Beech, D. J. (2001) Melanoma in African-Americans: trends in
20             biological behavior and clinical characteristics over two decades. J. Surg. Oncol. 78: 10-16.
21       Berntsen, T. K.; Isaksen, I. S. A. (1999) Effects of lightning and convection on changes in tropospheric ozone due to
22             NOX emissions from aircraft. Tellus 5IB: 766-788.
23       Berntsen, T. K.; Isaksen, I. S. A.; Myhre, G.; Fuglestvedt, J. S.; Stordal, F.; Larsen, T. A.; Freckleton, R. S.; Shine,
24             K. P. (1997) Effects of anthropogenic emissions on tropospheric ozone and its radiative forcing. J. Geophys.
25             Res. [Atmos.] 102: 28,101-28,126.
26       Berntsen, T. K.; Myhre, G.; Stordal, F.; Isaksen, I. S. A. (2000) Time evolution of tropospheric ozone and its
27             radiative forcing. J. Geophys. Res. [Atmos.] 105: 8915-8930.
28       Berret, J.; Liardet, S.; Scaletta, C.; Panizzon, R.; Hohlfeld, P.; Applegate, L. A. (2002) Use of sunscreens in families
29             living in Switzerland. Dermatology 204: 202-208.
30       Black, W. C.; Goldhahn, R. T., Jr.; Wiggins, C. (1987) Melanoma within a southwestern Hispanic population.
31             Arch. Dermatol. 123: 1331-1334.
32       Brasseur, G. P.; Kiehl,  J. T.; Mtiller, J.-F.; Schneider, T.; Granier, C.; Tie, X.-X.; Hauglustaine, D. (1998) Past and
33             future changes in global tropospheric ozone: impact on radiative forcing. Geophys. Res. Lett. 25: 3807-3810.
34       Breuer-McHam, J.; Simpson, E.; Dougherty, L; Bonkobara, M.; Ariizumi, K.; Lewis, D. E.; Dawson, D. B.;
3 5             Duvic, M.;  Cruz, P.O., Jr. (2001) Activation of HIV in human skin by ultraviolet B radiation and its
36             inhibition by NFKB blocking agents. Photochem.  Photobiol. 74: 805-810.
37       Briihl, C.; Crutzen, P. J. (1988) On the disproportionate role of tropospheric ozone as a filter against solar UV-B
38             radiation. Geophys. Res. Lett. 16: 703-706.
39       Chalita, S.; Hauglustaine, D. A.;  Le Treut, H.; Mtiller, J.  F.; Penkett,  S., eds. (1996) Radiative forcing due to
40             increased tropospheric ozone concentrations. Atmos. Environ. 30: 1641-1646.
41       Clydesdale, G. J.;  Dandie, G. W.; Muller, H. K. (2001) Ultraviolet light induced injury: immunological and
42             inflammatory effects. Immunol. Cell Biol. 79:  547-568.
43       Cockell, C. S. (2001) A photobiological history of Earth. In: Cockell, C. S.; Blaustein, A. R., eds. UV radiation flux
44             into ecosystems. New York, NY: Springer-Verlag.
45       Cutchis, P. (1974) Stratospheric ozone depletion and solar ultraviolet radiation on Earth. Science (Washington, DC)
46             184: 13-19.
47       Davis, K. J.; Cokkinides, V. E.; Weinstock, M. A.; O'Connell, M. C.; Wingo, P. A. (2002) Summer sunburn and sun
48             exposure among US youths ages 11 to 18: national prevalence and associated factors. Pediatrics 110: 27-35.
49       De Fabo, E. C.; Noonan, F. P.; Fears, T.; Merlino, G. (2004) Ultraviolet B but not ultraviolet A radiation initiates
50             melanoma.  Cancer Res. 64: 6372-6376.
51       De Gruijl, F. R. (1995) Action spectrum for photocarcinogenesis.  Recent Results Cancer Res. 139: 21-30.
52       De Gruijl, F. R. (1999) Skin cancer and solar UV radiation. Eur. J. Cancer 35: 2003-2009.
53       De Gruijl, F. R. (2002) Photocarcinogenesis: UVA vs. UVB radiation. Skin Pharmacol. Appl. Skin Physiol.
54             15:316-320.
         January 2005                                   10-50        DRAFT-DO NOT QUOTE OR CITE

-------
  1       De Gruijl, F. R.; Henricus, J. C. M.; Sterenborg, H.; Forbes, P. D.; Davies, R. E>; Cole, C.; Kelfkens, G.;
  2             Van Weelden, H.; Slaper, H.; VanderLeun, J. C. (1993) Wavelength dependence of skin cancer induction by
  3             ultraviolet irradiation of albino hairless mice. Cancer Res. 53: 53-60.
  4       Diepgen, T. L.; Mahler, V. (2002) The epidemiology of skin cancer. Br. J. Dermatol. 146(suppl. 61): 1-6.
  5       Diffey, B. L. (1990) Human exposure to ultraviolet radiation. Semin. Dermatol. 9: 2-10.
  6       Diffey, B. L. (1994) Observed and predicted minimal erythema doses: a comparative study. Photochem. Photobiol.
  7             60:380-382.
  8       Diffey, B. L.; Gibson, C. J.; Haylock, R.; McKinlay, A. F. (1996) Outdoor ultraviolet exposure of children and
  9             adolescents. Br. J. Dermatol.  134: 1030-1034.
10       Evans, R. D.; Kopf, A. W.; Lew, R.  A.; Rigel, D. S.; Bart, R. S.; Friedman, R. J.; Rivers, J. K. (1988) Risk factors
11             for the development of malignant melanoma--!: review of case-control studies. J. Dermatol. Surg. Oncol.
12             14:393-408.
13       Evans, W. F. J.; Ferguson, C. R.; Puckrin, E.  (2004) Radiative forcing by tropospheric ozone. Bull. Am. Meteorol.
14             Soc. 85: 347.
15       Fears, T. R.; Bird, C. C.; Guerry, D., IV; Sagebiel, R. W.; Gail, M. H.; Elder, D. E.; Halpern, A.; Holly, E. A.;
16             Hartge, P.; Tucker M. A. (2002) Average midrange ultraviolet radiation flux and time outdoors predict
17             melanoma risk. Cancer Res. 62: 3992-3996.
18       Fiore, A. M.; Jacob, D. J.; Field, B. D.; Streets,  D. G.; Fernandes, S. D.; Jang, C. (2002) Linking ozone pollution and
19             climate change: the case for controlling methane. Geophys. Res. Lett.  29(19):  10.1029/2002GL015601.
20       Fishman, J.; Wozniak, A. E.; Creilson, J. K. (2003) Global distribution of tropospheric  ozone from satellite
21             measurements using the empirically corrected tropospheric ozone residual technique: identification of the
22             regional aspects of air pollution. Atmos.  Chem. Phys. 3: 893-907.
23       Fligge, M.; Solanki, S. K. (2000) Reconstruction of past solar irradiance.  Space Sci. Rev. 94: 127-138.
24       Frederick, J. E.; Koob, A. E.; Weatherhead, E. C. (1993) Empirical studies of tropospheric transmission in the
25             ultraviolet: roadband measurements. J. Appl. Meteorol. 32: 1883-1892.
26       Freedman, D. M.; Dosemeci, M.; McGlynn, K.  (2002) Sunlight and mortality from breast, ovarian, colon,  prostate,
27             and non-melanoma skin cancer: a composite death certificate based case-control study. Occup. Environ. Med.
28             59:257-262.
29       Gallagher, R. P.; Hill, G. B.; Bajdik, C. D.; Coldman, A. J.; Fincham, S.; McLean, D. I.; Threlfall, W. J. (1995a)
3 0             Sunlight exposure, pigmentation factors and risk of nonmelanocytic skin cancer. II. Squamous cell carcinoma.
31             Arch. Dermatol. 131:  164-169.
32       Gallagher, R. P.; Hill, G. B.; Bajdik, C. D.; Fincham, S.; Coldman, A. J.; McLean, D. I.; Threlfall, W. J. (1995b)
33             Sunlight exposure, pigmentary factors, and risk of nonmelanocytic skin cancer. I. Basal cell carcinoma.
34             Arch. Dermatol. 131:  157-163.
35       Garland, F. C.; Garland, C. F.; Gorham, E. D.; Young, J. F. (1990) Geographic variation in breast cancer mortality in
36             the United States: a hypothesis involving exposure to solar radiation. Prev. Med. 19: 614-622.
37       Garssen, J.; Van Loveren, H. (2001) Effects of ultraviolet exposure  on the immune system. Crit. Rev. Immunol.
38             21:359-397.
39       Gauss, M.; Myhre, G.; Pitari, G.; Prather, M.  J.; Isaksen, I. S. A.; Berntsen, T. K.; Brasseur, G. P.; Dentener, F. J.;
40             Derwent, R. G.; Hauglustaine, D. A.; Horowitz, L. W.; Jacob, D. J.; Johnson, M.; Law, K. S.; Mickley, L.  J.;
41             Mtiller, J.-F.;  Plantevin, P. H.; Pyle, J.  A.; Rogers, H. L.; Stevenson, D. S.; Sundet, J. K.; Van Weele, M.;
42             Wild, O. (2003) Radiative forcing in the  21st century due to ozone changes in the troposphere and the lower
43             stratosphere. J. Geophys. Res. (Atmos.) 108(D9): 10.1029/2002JD002624.
44       Gies, P.; Wright,  J. (2003) Measured solar ultraviolet radiation exposures of outdoor workers in Queensland in the
45             building and construction industry. Photochem. Photobiol. 68: 78-83.
46       Gies, P.; Roy, C.; Toomey, S.; MacLennan, R.;  Watson, M.  (1998) Solar UVR exposures of primary school children
47             at three locations in Queensland. Photochem. Photobiol. 68: 78-83.
48       Gloster, H. M., Jr.; Brodland, D. G. (1996) The epidemiology of skin cancer. Dermatol. Surg. 22: 217-226.
49       Godar, D. E. (2001)  UV doses of American children and adolescents. Photochem. Photobiol. 74: 787-793.
50       Godar, D. E.; Wengraitis, S. P.;  Shreffler, J.; Sliney, D. H. (2001) UV doses of Americans. Photochem. Photobiol.
51             73:621-629.
52       Godar, D. E.; Urbach, F.; Gasparro, F. P.; Van Der Leun, J. C. (2003) UV doses of young adults. Photochem.
53             Photobiol. 77: 453-457.
54       Gorham, E. D.; Garland, F. C.; Garland, C. F. (1990) Sunlight and breast  cancer incidence in the USSR. Int. J.
55             Epidemiol. 19: 820-824.
56       Granstein, R. D.; Matsui, M. S.  (2004) UV radiation-induced immunosuppression and skin cancer. Cutis
57             74(suppl. 5): 4-9.


         January 2005                                   10-51        DRAFT-DO NOT QUOTE OR CITE

-------
  1       Grant, W. B. (2002a) An estimate of premature cancer mortality in the U.S. due to inadequate doses of solar
  2             ultraviolet-B radiation. Cancer 94:  1867-1875.
  3       Grant, W. B. (2002b) An ecologic study of dietary and solar ultraviolet-B links to breast carcinoma mortality rates.
  4             Cancer 94: 272-281.
  5       Great Lakes Regional Assessment Group  (GLRAG). (2000) Preparing for a changing climate: the potential
  6             consequences of climate variability and change. Great Lakes overview. Washington, DC: U.S. Environmental
  7             Protection Agency;  Office of Research and Development; U.S. Global Change Research Program (USGCRP)
  8             and Ann Arbor, MI: University of Michigan; Atmospheric, Oceanic and Space Sciences Department.
  9             Available: http://www.geo.msu.edu/glra/assessment/assessment.html [17 April 2002].
10       Green, A.; Battistutta, D.; Hart, V.; Leslie, D.; Weedon, D. (1996) Skin cancer in a subtropical Australian
11             population: incidence and lack of association with occupation. The Nambour Study Group. Am. J. Epidemiol.
12             144: 1034-1040.
13       Hall, H. I.; Miller, D. R.; Rogers, J. D.; Bewerse, B. J. (1999) Update on the incidence and mortality from melanoma
14             in the United States. J. Am. Acad. Dermatol. 40: 35-42.
15       Halliday, G. M.; Byrne,  S. N.; Kuchel, J. M.; Poon, T. S.; Barnetson, R. S. (2004) The suppression of immunity by
16             ultraviolet radiation: UVA, nitric oxide and DNA damage. Photochem. Photobiol. Sci. 3: 736-740.
17       Hanchette, C. L.; Schwartz, G. G. (1992)  Geographic patterns of prostate cancer mortality. Cancer 70: 2861-2869.
18       Harrison, G. I.; Young, A.  R. (2002) Ultraviolet radiation-induced erythema in human skin. Methods (San Diego,
19             CA, U.S.) 28: 14-19.
20       Hartge, P.; Devesa, S. S.; Grauman, D.; Fears, T. R.; Fraumeni, J. F., Jr. (1996) Non-Hodgkin's lymphoma and
21             sunlight. J. Natl. Cancer Inst. 88: 298-300.
22       Hartmann, D. L. (1994)  Global physical climatology. New York, NY: Academic Press. (International geophysics:
23             v. 56).
24       Hauglustaine, D. A.; Brasseur, G. P.; Walters, S.; Rasch, P. J.; Mtiller, J.-F.; Emmons, L. K.; Carroll, M. A. (1998)
25             MOZART, a global chemical transport model for ozone and related chemical tracers 2. model results and
26             evaluation. J. Geophys. Res. (Atmos.) 103: 28291-28335.
27       Haywood,  J. M.; Schwarzkopf, M. D.; Ramaswamy, V. (1998) Estimates of radiative forcing due to modeled
28             increases in tropospheric ozone. J.  Geophys. Res. [Atmos.] 103: 16,999-17,007.
29       Henriksen, M.; Na, R.; Agren, M. S.; Wulf, H. C. (2004) Minimal erythema dose after umltiple UV exposures
30             depends on pre-exposure skin pigmentation. Photodermatol. Photoimmunol. Photomed. 20:  163-169.
31       Hildesheim, J.; Fornace, A. J., Jr. (2004) The dark side of light: the damaging effects of UV rays and the protective
32             efforts of MAP kinase signaling in the epidermis. DNA Repair 3: 567-580.
33       Hockwin, O.; Kojima, M.; Sakamoto, Y.;  Wegener, A.; Shui, Y. B.; Sasaki, K. (1999) UV damage to the eye lens:
34             further results from animal model studies: a review. J. Epidemiol. 9(suppl. 6): S39-S47.
3 5       Holick, M. F. (2004) Sunlight and vitamin D for bone health and prevention of autoimmune diseases, cancers, and
36             cardiovascular disease. Am. J. Clin. Nutr. 80(6  suppl.): 1678S-1688S.
37       Hu, S.; Ma, F.; Collado-Mesa, F.; Kirsner, R. S. (2004) UV radiation, latitude, and melanoma in US Hispanics and
38             blacks. Arch. Dermatol. 140: 819-824.
39       Ichihashi, M.; Ueda, M.; Budiyanto, A.; Bito, T.; Oka, M.; Fukunaga, M.; Tsuru, K.; Horikawa, T. (2003)
40             UV-induced skin damage. Toxicology 189: 21-39.
41       Intergovernmental Panel on Climate Change. (1996) Climate change 1995: the science of climate change. Summary
42             for policymakers  and technical summary of the working group I report. Presented at: the fifth session of the
43             IPCC Working Group I; November; Madrid, Spain. Geneva,  Switzerland: World Meteorological Association.
44       Intergovernmental Panel on Climate Change (IPCC). (2001a) Climate change 2001: the scientific basis. Contribution
45             of working group I to  the third assessment report of the Intergovernmental Panel on Climate Change.
46             Cambridge, United Kingdom: Cambridge University Press.
47       Intergovernmental Panel on Climate Change (IPCC). (200 Ib) Climate change 2001: impacts, adaptation, and
48             vulnerability. Contribution of working group II to the third assessment report of the Intergovernmental Panel
49             on Climate Change. Cambridge, United Kingdom: Cambridge University Press.
50       International Agency for Research on Cancer. (1992) Solar and ultraviolet radiation. Lyon, France: International
51             Agency for Research on Cancer. (IARC monographs on the evaluation of carcinogenic risks to humans:
52             v. 55).
53       International Commission on Non-Ionizing Radiation Protection. (2004) Guidelines on limits of exposure to
54             ultraviolet radiation of wavelengths between 180 nm and  400 nm (incoherent optical radiation). Health Phys.
55             87:  171-186.
56       Javitt, J. C.; Taylor, H. R. (1994-95) Cataract and latitude.  Doc.  Ophthalmol. 88: 307-325.
         January 2005                                  10-52        DRAFT-DO NOT QUOTE OR CITE

-------
  1       Jemal, A.; Tiwari, R. C.; Murray, T.; Ghafoor, A.; Samuels, A.; Ward, E.; Feuer, E. I; Thun, M. J. (2004) Cancer
  2             statistics, 2004. CA-Cancer J. Clin. 54: 8-29.
  3       John, E. M.; Schwartz, G. G.; Dreon, D. M.; Koo, J. (1999) Vitamin D and breast cancer risk: the NHANES I
  4             Epidemiologic Follow-up Study, 1971-1975 to 1992. Cancer Epidemiol. Biomarkers Prev. 8: 399-406.
  5       Kaidbey, K. H.; Kligman, A. M. (1981) Cumulative effects from repeated exposures to ultraviolet radiation.
  6             J. Invest. Dermatol. 76: 352-355.
  7       Kerr, J. B.; McElroy, C. T. (1993) Evidence for large upward trends of ultraviolet-B radiation linked to ozone
  8             depletion. Science (Washington, DC) 262: 1032-1034.
  9       Kiehl, J. T.; Schneider, T. L.; Portmann, R. W.; Solomon, S. (1999) Climate forcing due to tropospheric and
10             stratospheric ozone. J. Geophys. Res. [Atmos.] 104: 31,239-31,254.
11       Kimlin, M.  G.; Parisi, A. V.; Wong, J.  C. (1998) Quantification of personal solar UV exposure of outdoor workers,
12             indoor workers and adolescents  at two locations in Southeast Queensland. Photodermatol. Photoimmunol.
13             Photomed. 14:7-11.
14       Kimlin, M.  G.; Wong, J. C.; Parisi, A. V. (1998) Simultaneous comparison of the personal UV exposure of two
15             human groups at different altitudes. Health Phys. 74: 429-434.
16       Kollias, N.; Stamatas, G. N.; Youn, J. I. (2001) Suppression of UVB-induced cutaneous erythema by aprevious
17             UVB exposure. Photochem. Photobiol. 74: 471-476.
18       Koloutsou-Vakakis, S.; Helmis, C. G.;  Assimakopoulos, V.; Glisten, H. (2001) Middle and lower troposphere
19             aerosol characteristics and ozone concentrations over northwestern Greece during STAAARTE 1997.
20             Atmos. Environ. 35: 1517-1526.
21       Kraemer, K. H. (1997) Sunlight and skin cancer: another link revealed. Proc. Natl. Acad. Sci. U. S. A. 94: 11-14.
22       Kricker, A.; Armstrong, B. K.; English, D. R. (1994) Sun exposure and non-melanocytic skin cancer. Cancer Causes
23             Control  5: 367-392.
24       Krol, M.; Van Leeuwen, P. J.; Lelieveld, J. (1998) Global OH trend inferred from methylchloroform measurements.
25             J. Geophys. Res. [Atmos.] 103:  10,697-10,711.
26       Krotkov, N. A.; Bhartia, P. K.; Herman, J. R.; Fioletov, V; Kerr, J. (1998) Satellite estimation of spectral surface
27             UV irradiance in the presence of tropospheric aerosols 1. cloud-free case. J. Geophys. Res.  (Atmos.)
28             103:  8779-8793.
29       Lefkowitz, E. S.; Garland, C. F. (1994) Sunlight, vitamin D, and ovarian cancer mortality rates in U.S. women.
30             Int. J. Epidemiol. 23: 1133-1136.
31       Lelieveld, J.; Dentener, F. J. (2000) What controls tropospheric ozone? J. Geophys. Res. [Atmos.] 105: 3531-3551.
32       Lens, M. B.; Dawes, M. (2004) Global perspectives of contemporary epidemiological trends of cutaneous malignant
33             melanoma. Br. J. Dermatol. 150: 179-185.
34       Levy, H., II; Kasibhatla, P. S.; Moxim, W. J.; Klonecki, A. A.; Hirsch, A. I.; Oltmans, S. J.; Chameides, W. L.
35             (1997) The global impact of human activity on tropospheric ozone. Geophys. Res. Lett.  24: 791-794.
36       Liao, H.; Seinfeld, J. H.; Adams, P. J.;  Mickley, L. J. (2004) Global radiative forcing of coupled tropospheric ozone
37             and aerosols in a unified general circulation model. J.  Geophys. Res. [Atmos.] 109(D16):
38             10.1029/2003 JD004456.
39       Logan, J. A. (1999) An analysis of ozonesonde data for the troposphere: recommendations for testing 3-D models
40             and development of a gridded climatology for tropospheric ozone. J. Geophys. Res. 104: 16,115-16,149.
41       Longstreth, J.; De Gruijl, F. R.; Kripke, M. L.; Abseck, S.; Arnold, F.; Slaper, H. I.; Velders, G.; Takizawa, Y.;
42             Van Der Leun, J. C. (1998) Health risks. J. Photochem. Photobiol. B 46: 20-39.
43       Lovato, C. Y.;  Shoveller, J. A.; Peters,  L.; Rivers, J. K. (1998a) Canadian National Survey on Sun Exposure &
44             Protective Behaviours: parent's reports on children. Cancer Prev. Control.  2:  123-128.
45       Lovato, C. Y.;  Shoveller, J. A.; Peters,  L.; Rivers, J. K. (1998b) Canadian National Survey on Sun Exposure &
46             Protective Behaviours: youth at  leisure. Cancer Prev. Control 2: 117-122.
47       Madronich, S.; De Gruijl, F. (1993) Skin cancer and UV radiation. Nature (London, UK) 366(6450): 23.
48       Madronich, S.; De Gruijl, F. (1994) Reply to "Kricker, A.; Armstrong, B. K.; McMichael, A. J. (1994) Skin cancer
49             and ultraviolet. Nature (London, UK) 368: 594." Nature (London, UK) 368(6472): 594.
50       Matsumura, Y.; Ananthaswamy, H. N.  (2004) Toxic effects of ultraviolet radiation on the  skin. Toxicol. Appl.
51             Pharmacol. 195: 298-308.
52       McCarty, C. A.; Nanjan, M. B.; Taylor, H. R. (2000) Attributable risk estimates for cataract to prioritize medical and
53             public health action. Invest. Ophthalmol. Vis. Sci. 41: 3720-3725.
54       McCarty, C. A.; Taylor, H. R. (2002) A review of the epidemiologic evidence linking ultraviolet radiation and
55             cataracts. Dev. Ophthalmol. 35:  21-31.
         January 2005                                   10-53         DRAFT-DO NOT QUOTE OR CITE

-------
  1       McKenzie, R.; Smale, D.; Bodeker, G.; Claude, H. (2003) Ozone profile differences between Europe and New
  2             Zealand: effects on surface UV irradiance and its estimation from satellite sensors. J. Geophys. Res. [Atmos.]
  3             108(D6): 10.1029/2002JD002770.
  4       McKinlay, A. F.; Diffey, B. L. (1987) A reference action spectrum for ultraviolet induced erythema in human skin.
  5             In: Passchier, W. F.; Bosnijakovic, B. F. M, eds. Human exposure to ultraviolet radiation: risks and
  6             regulations. Amsterdam, The Netherlands: Elsevier Science Publishers; pp. 83-87.
  7       Merriam, J. C.; Lofgren, S.; Michael, R.; Soderberg, P.; Dillon, J.; Zheng, L.; Ayala, M. (2000) An action spectrum
  8             for UV-B radiation and the rat lens. Invest. Ophthalmol. Vis. Sci. 41:  2642-2647.
  9       Mickley, L. J.; Murti, P. P.; Jacob, D. J.; Logan, J. A.; Koch, D. M.; Rind, D. (1999) Radiative forcing from
10             tropospheric ozone calculated with a unified chemistry-climate model. J. Geophys. Res. [Atmos.]
11             104:30,153-30,172.
12       Mickley, L. J.; Jacob, D. J.; Field, B. D.; Rind, D. (2004) Climate response to the increase in tropospheric ozone
13             since preindustrial times: a comparison between ozone and equivalent CO2 forcings. J. Geophys. Res.
14             [Atmos.] 109(05): 10.1029/2003 JD003653.
15       Mid-Atlantic Regional Assessment (MARA) Team. (2000) Preparing for a changing climate:  the potential
16             consequences of climate variability and change. Mid-Atlantic overview. Washington, DC: U.S.
17             Environmental Protection Agency; Office of Research and Development; U.S. Global Change Research
18             Program (USGCRP) and University Park, PA: Pennsylvania State University. Available:
19             http://www.usgcrp.gov/usgcrp/nacc/midatlantic.htm [17 April 2002].
20       Moehrle, M. (2001) Ultraviolet exposure in the Ironman triathlon. Med. Sci. Sports Exercise 33: 1385-1386.
21       Moehrle, M.; Heinrich, L.; Schmid, A.; Garbe, C. (2000) Extreme UV exposure of professional cyclists.
22             Dermatology 201: 44-45.
23       Moise, A.  F.; Buttner, P. G.; Harrison, S. L. (1999) Sun exposure at school. Photochem. Photobiol. 70: 269-274.
24       National Assessment Synthesis Team (NAST). (2000) Climate change impacts on the United States: the potential
25             consequences of climate variability and change. Overview. Washington, DC: U.S. Global Change Research
26             Program. Available at: http://www.usgcrp.gov/usgcrp/Library/nationalassessment/overview.htm
27             (17 January 2003).
28       National Research Council. (2005) Radiative forcing of climate change: expanding the concept and addressing
29             uncertainties. Washington, DC: The National Academies Press.
30       New England Regional Assessment Group (NERAG). (2001) Preparing for a changing climate: the potential
31             consequences of climate variability and change. New England regional overview. Washington, DC: U.S.
32             Environmental Protection Agency; Office of Research and Development; U.S. Global Change Research
3 3             Program (USGCRP) and Durham, NH: University of New Hampshire; Institute for the Study of Earth,
34             Oceans, and Space. Available: http://www.necci.sr.unh.edu/2001-NERA-report.html [17 April 2002].
3 5       Nole, G.; Johnson, A. W. (2004) An analysis of cumulative lifetime solar ultraviolet radiation exposure and the
36             benefits of daily sun protection. Dermatol. Ther.  17(suppl. 1): 57-62.
37       Noonan, F. P.; Muller,  H. K.; Fears, T. R.; Kusewitt, D. F.; Johnson, T. M.; De Fabo, E. D. (2003) Mice with
3 8             genetically determined high susceptibility to ultraviolet (UV)-induced immunosuppression show enhanced
39             UV carcinogenesis. J. Invest. Dermatol. 121: 1175-1181.
40       Norval, M.; Garssen, J.; Van Loveren, H.; El-Ghorr, A. A. (1999) UV-induced changes in the immune response to
41             microbial infections in human subjects and animal models. J. Epidemiol. 9(suppl. 6): S84-S92.
42       Okada, S.; Weatherhead, E.; Targoff, I. N.; Wesley, R.; Miller, F.  W.; International Myositis Collaborative Study
43             Group. (2003) Global surface ultraviolet radiation intensity may modulate the clinical and imunologic
44             expression of autoimmune muscle disease. Arthritis Rheum. 48: 2285-2293.
45       Oriowo, O. M.; Cullen, A. P.; Chou, B. R.; Sivak, J. G.  (2001) Action spectrum and recovery for in vitro
46             UV-induced cataract using whole lenses. Invest. Ophthalmol. Vis. Sci. 42: 2596-2602.
47       Parisi, A. V.; Willey, A.; Kimlin, M. G.; Wong, J. C. (1999) Penetration of solar erythema! UV radiation in the
48             shade of two common Australian trees. Health Phys. 76: 682-686.
49       Patz, J. A.; Engelberg,  D.; Last, J. (2000a) The effects of changing weather on public health. Annu. Rev. Public
50             Health 21: 271-307.
         January 2005                                   10-54         DRAFT-DO NOT QUOTE OR CITE

-------
  1       Patz, J. A.; McGeehin, M. A.; Bernard, S. M; Ebi, K. L.; Epstein, P. R.; Grambsch, A.; Gubler, D. I; Reiter, P.;
  2             Romieu, L; Rose, J. B.; Samet, J. M.; Trtanj, J. (2000b) The potential health impacts of climate variability and
  3             change for the United States: executive summary of the report of the health sector of the U.S. national
  4             assessment. Environ. Health Perspect. 108: 367-376.
  5       Persson, A. E.; Edstrom, D. W.; Backvall, H.; Lundeberg, J.; Ponten, F.; Ros, A. M.; Williams, C. (2002) The
  6             mutagenic effect of ultraviolet-A1 on human skin demonstrated by sequencing the p53 gene in single
  7             keratinocytes. Photodermatol. Photoimmunol. Photomed. 18: 287-293.
  8       Pfister, H. (2003) Human papillomavirus and skin cancer. J. Natl.  Cancer Inst. Monogr. 31: 52-56.
  9       Pitts, D. G.  (1993) Ocular effects of radiant energy. In: Pitts, D. G.; Kleinstein, R. N., eds. Environmental Vision.
10             Stoneham, MA: Butterworth-Heinemann, pp. 151-220.
11       Prinn, R. G.; Weiss, R. F.; Miller, B. R.; Huang, J.; Alyea, F. N.; Cunnold, D. M.; Fraser, P. J.; Hartley, D. E.;
12             Simmonds, P. G. (1995) Atmospheric trends and lifetime of CH3CC13 and global OH concentrations. Science
13             (Washington, DC) 269: 187-192.
14       Rigel, D. S.; Rigel, E. G.; Rigel, A. C. (1999) Effects of altitude and latitude on ambient UVB radiation. J. Am.
15             Acad. Dermatol. 40: 114-116.
16       Rivers, J. K. (2004) Is there more than one road to melanoma? Lancet 363: 728-730.
17       Robinson, E. S.; Hill, R. H., Jr.; Kripke, M. L.; Setlow, R. B. (2000) The Monodelphis melanoma model: initial
18             report on large ultraviolet A exposures of suckling young. Photochem. Photobiol. 71: 743-746.
19       Roelofs, G. J.; Lelieveld, J.; Van Borland, R. (1997) A three-dimensional chemistry/general circulation model
20             simulation of anthropogenically derived ozone in the troposphere and its radiative climate forcing.
21             J. Geophys. Res. [Atmos.] 102: 23,389-23,401.
22       Rosenthal, F. S.; Phoon, C.; Bakalian, A. E.; Taylor, H. R. (1988) The ocular dose of ultraviolet radiation to outdoor
23             workers. Invest. Ophthalmol. Vis. Sci. 29: 649-656.
24       Rozema, J.; Manetas, Y.; Bjorn, L. O., eds. (2001) Responses of plants to UV-B radiation. Dordrecht, The
25             Netherlands: Kluwer Academic. [Advances in Vegetation Science, v. 18].
26       Runger, T. M.; Moller, K.; Jung, T.; Dekant, B. (2000) DNA damage formation, DNA repair, and survival after
27             exposure  of DNA repair-proficient and nucleotide excision repair-deficient human lymphoblasts to UVA1
28             and UVB. Int. J. Radiat. Biol. 76: 789-797.
29       Salby, M. L. (1996) Fundamentals of atmospheric physics. New York, NY: Academic Press. (The International
30             Geophysics Series, Monographs and Textbooks, v. 61).
31       Sarasin, A.  (1999) The molecular pathways of ultraviolet-induced carcinogenesis. Mutat. Res. 428: 5-10.
32       Sarkar, A. K. (2004) An evaluation of UV protection imparted by  cotton fabrics dyed with natural colorants. BMC
33             Dermatol. 4: 15.
34       Schenker, M. B.; Orenstein, M. R.; Samuels, S. J. (2002) Use of protective equipment among California farmers.
35             Am.  J. Ind. Med. 42: 455-464.
36       Selgrade, M. K.; Repacholi, M. H.; Koren, H. S. (1997) Ultraviolet radiation-induced immune modulation: potential
37             consequences for infectious, allergic, and autoimmune disease. Environ. Health Perspect.  105: 332-334.
38       Selgrade, M. K.; Smith, M. V; Oberhelman-Bragg, L. J.; LeVee, G. J.; Koren, H. S.; Cooper, K. D. (2001) Dose
3 9             response for UV-induced immune suppression in people of color: differences based on erythema! reactivity
40             rather than skin pigmentation. Photochem. Photobiol. 74: 88-95.
41       Setlow, R. B.; Grist, E.; Thompson, K.; Woodhead, A. D. (1993) Wavelengths effective in induction of malignant
42             melanoma. Proc. Natl. Acad. Sci. U. S. A. 90: 6666.6670.
43       Shindell, D. T.; Faluvegi, G.; Bell, N. (2003) Preindustrial-to-present-day radiative forcing by tropospheric ozone
44             from improved simulations with the GISS chemistry-climate GCM. Atmos. Chem. Phys. 3: 1675-1702.
45       Shoveller, J. A.; Lovato, C. Y.; Peters, L.; Rivers, J. K. (1998) Canadian National Survey on Sun Exposure &
46             Protective Behaviours: adults at leisure. Cancer Prev. Control 2: 111-116.
47       Slaper, H.; Velders, G. J. M.; Daniel, J. S.; de Gruijl, F. R.; Van derLeun, J. C. (1996) Estimates of ozone depletion
48             and skin cancer incidence to examine the Vienna Convention achievements. Nature (London, U.K.)
49             384:256-258.
50       Stern, R. S.; Liebman, E. J.; Vakeva, L. (1998) Oral psoralen and ultraviolet-A light (PUVA) treatment of psoriasis
51             and persistent risk of nonmelanoma skin cancer. PUVA Follow-up Study. J. Natl. Cancer Inst. 90: 1278-1284.
52       Stevenson, D. S.; Johnson, C. E.; Collins, W. J.; Derwent, R. G.; Shine, K. P.; Edwards, J. M. (1998) Evolution of
53             tropospheric ozone radiative forcing.  Geophys. Res. Lett. 25: 3819-3822.
54       Stevenson, D. S.; Johnson, C. E.; Collins, W. J.; Derwent, R. G.; Edwards, J. M. (2000) Future estimates of
55             tropospheric ozone radiative forcing and methane turnover - the impact of climate change. Geophys. Res.
56             Lett. 27: 2073-2076.
57       Studzinski,  G. P.; Moore, D. C. (1995) Sunlight-can it prevent as well as cause cancer? Cancer Res. 55: 4014-4022.


         January 2005                                   10-55         DRAFT-DO NOT QUOTE OR CITE

-------
  1       Swetter, S. M. (2003) Dermatological perspectives of malignant melanoma. Surg. Clin. N. Am. 83: 77-95.
  2       Taylor, H. R.; West, S. K.; Rosenthal, F. S.; Munoz, B.; Newland, H. S.; Abbey, H.; Emmett, E. A. (1988) Effect of
  3             ultraviolet radiation on cataract formation. N. Engl. J. Med. 319:  1429-1433.
  4       Thieden, E.; Agren, M. S.; Wulf, H. C. (2001) Solar UVR exposures of indoor workers in a working and a holiday
  5             period assessed by personal dosimeters and sun exposure diaries. Photodermatol. Photoimmunol. Photomed.
  6             17:249-255.
  7       Thieden, E.; Philipsen, P. A.; Heydenreich, J.; Wulf, H. C. (2004a) UV radiation exposure related to age, sex,
  8             occupation, and sun behavior based on time-stamped personal dosimeter readings. Arch. Dermatol. 140:
  9             197-203.
10       Thieden, E.; Philipsen, P. A.; Sandby-Moller, J.; Heydenreich, J.; Wulf, H. C. (2004b) Proportion of lifetime UV
11             dose received by children, teenagers and adults based on time-stamped personal dosimetry. J. Invest.
12             Dermatol. 123: 1147-1150.
13       Tomescu, D.; Kavanagh, G.; Ha, T.; Campbell, H.; Melton, D. W. (2001) Nucleotide excision repair geneXPD
14             polymorphisms and genetic predisposition to melanoma. Carcinogenesis 22: 403-408.
15       U.S. Environmental Protection Agency. (1987) Assessing the risks of trace gases that can modify the stratosphere,
16             v. 1-8. Washington, DC: Office of Air and Radiation; report no. EPA 400/1-87/001A-H.
17       U.S. Environmental Protection Agency. (2004) Air quality criteria for particulate matter. Research Triangle Park,
18             NC: National Center for Environmental Assessment; report no. EPA/600/P-99/002aF-bF. 2v. Available:
19             http://cfpub.epa.gov/ncea/ [9 November, 2004].
20       United Nations Environment Programme (UNEP). (1986) Report of the  international conference on the assessment
21             of the role of carbon dioxide and of other greenhouse gases in climate variations and associated impacts;
22             October 1985; Villach, Austria. Geneva, Switzerland: World Meteorological Organization; WMO no. 661.
23       Van Der Leun, J. C.; Takiyawa, Y.; Longstreth, J. D. (1989) Human health. In: Van der Leun, J. C.; Trevini, M.;
24             Worress,  R. C., eds. Environmental Effects Panel Report. Nairobi, Kenya: United Nations Environment
25             Programme, pp. 11-24. Available: http://www.ciesin.org/docs/001-538/001-538.html.
26       Van Dorland, R.; Dentener, F. J.; Lelieveld, J. (1997) Radiative forcing  due to tropospheric ozone and sulfate
27             aerosols.  J. Geophys. Res. [Atmos.] 102: 28,079-28,100.
28       Veiered, M. B.;  Weiderpass, E.; Thorn, M.; Hansson, J.; Lund, E.; Armstrong, B.; Adami, H.-O. (2003)
29             A prospective study of pigmentation, sun exposure, and risk of cutaneous malignant melanoma in women.
30             J. Natl. Cancer Inst. 95: 1530-1538.
31       Vishvakarman, D.; Wong, J. C.; Boreham, B. W. (2001) Annual occupational exposure to ultraviolet radiation in
32             central Queensland. Health Phys. 81: 536-544.
33       Vitasa, B. C.; Taylor, H. R.; Strickland, P. T.; Rosenthal, F. S.; West, S.; Abbey, H.; Ng, S. K.; Munoz, B.; Emmett,
34             E. A. (1990) Association on nonmelanoma skin cancer and actinic keratosis with cumulative solar ultraviolet
35             exposure  in Maryland watermen. Cancer 65:  2811-2817.
36       Waterston, K.; Nay smith, L.; Rees. J. L. (2004) Physiological variation in the erythemal response to ultraviolet
37             radiation  and photoadaptation. J. Invest. Dermatol. 123: 958-964.
3 8       Wayne, R. P. (2000) Chemistry of Atmospheres: an introduction to the chemistry of the atmospheres of Earth, the
39             planets, and their satellites. 3rd ed. New York, NY: Oxford University Press, Inc.
40       Wei, Q.; Lee, J.  E.; Gershenwald, J. E.; Ross, M. I.; Mansfield, P. F.; Strom, S. S.; Wang, L. E.; Guo, Z.; Qiao, Y.;
41             Amos, C. I.; Spitz, M. R.; Duvic, M. (2003) Repair of UV light-induced DNA damage and risk of cutaneous
42             malignant melanoma. J. Natl. Cancer Inst. 95: 308-315.
43       Wendisch, M.; Mayer, B. (2003) Vertical distribution of spectral solar irradiance in the cloudless sky: a case study.
44             Geophys. Res. Lett. 30:  10.1029/2002GL016529.
45       Westerdahl, J.; Ingvar, C.; Masback, A.; Jonsson, N.; Olsson, H. (2000) Risk of cutaneous malignant melanoma in
46             relation to use of sunbeds: further evidence for UV-A carcinogenicity. Br. J. Cancer 82: 1593-1599.
47       Whiteman, D. C.; Valery, P.; McWhirter, W.; Green, A. C. (1997) Risk factors for childhood malanoma in
48             Queensland, Australia.  Int. J. Cancer 70: 26-31.
49       Whiteman, D. C.; Whiteman, C. A.; Green, A. C. (2001) Childhood sun exposure as a risk factor for melanoma:
50             a systematic review of epidemiologic studies. Cancer Causes Control 12: 69-82.
51       Wigley, T. M. L.; Smith, S. J.; Prather, M. J. (2002) Radiative forcing due to reactive gas emissions. J. Clim.
52             15:2690-2696.
53       Wong, J. C.; Airey, D. K.; Fleming, R. A. (1996) Annual reduction of solar UV exposure to the facial aea of outdoor
54             workers in southeast Queensland by wearing a hat. Photodermatol. Photoimmunol. Photomed. 12: 131-135.
55       Woodward, A.; Boffetta, P. (1997) Environmental exposure, social class, and cancer risk. In: Kogevinas, M.;
56             Pearce, N.; Susser, M.; Boffetta, P. Social inequalities and cancer. Lyon, France: International Agency for
57             Research on Cancer; pp. 361-367. (IARC scientific publications:  v. 138).


         January 2005                                    10-56       DRAFT-DO NOT QUOTE OR CITE

-------
 1       World Meteorological Organization. (1988) Developing policies for responding to climatic change: a summary of
 2             the discussions and recommendations of workshops; September-October 1987; Villach, Austria; and
 3             November 1987; Bellagio, Austria. Geneva, Switzerland: World Meteorological Organization; report no.
 4             WMO/TD; no. 225.  [World Climate Impact Programme series report no. WCIP-1].
 5       World Meteorological Organization (WMO). (2003) Scientific assessment of ozone depletion: 2003. Geneva,
 6             Switzerland: United Nations Environment Programme. [Global Ozone and Monitoring Project; report no. 47].
 7       Xenopoulos, M. A.; Schindler, D. W. (2001) Physical factors determining ultraviolet radiation flux into ecosystems.
 8             In: Cockell, C.; Blaustein, A. R. Ecosystems, evolution, and ultraviolet radiation. New York, NY: Springer;
 9             pp. 36-62.
10       Yarnal, B.; Kalkstein, L. S.; Scheraga, J. D. (2000) Mid-Atlantic regional assessment of climate change impacts.
11             Clim. Res. (CR Special 7) 14: 153-269.
12       Zeng, G.; Pyle, J. A. (2003) Changes in tropospheric ozone between 2000 and 2100 modeled in a chemistry-climate
13             model. Geophys. Res. Lett. 30: 10.1029/2002GLO16708.
14
15
         January 2005                                   10-57        DRAFT-DO NOT QUOTE OR CITE

-------
 i       11.  EFFECT OF OZONE ON MAN-MADE MATERIALS
 2          Ozone and other photochemical oxidants react with many of the economically important
 3     man-made materials, decreasing their useful life and aesthetic appearance.  Some of the
 4     materials known to be damaged by ozone include elastomers, fibers and dyes, and paints.  This
 5     section will provide a brief discussion on the effects of ozone on man-made materials including
 6     the damage mechanisms and, where possible, concentration-response relationships. Since only
 7     limited information has been published on the effects of ozone on materials, this section will
 8     provide a summary of information presented in the previous ozone criteria document (U.S.
 9     Environmental Protection Agency, 1996) and a more detailed discussion of studies published
10     since publication of the previous ozone criteria document. The reader is referred to the previous
11     ozone criteria document for a more detailed discussion of the earlier studies.
12
13     11.11.1  Mechanisms of Ozone Damage and Exposure-Response
14     Elastomer Cracking
15          The elastomeric compounds, natural rubber and synthetic polymers and copolymers of
16     butadiene, isoprene, and styrene, are particularly susceptible to even low levels of ozone.
17     Elastomeric compounds are long chain unsaturated organic molecules. Ozone damages these
18     compounds by breaking the molecular chain at the carbon-carbon double bond; a chain of three
19     oxygen atoms is added directly across the double bond, forming a five-membered ring structure
20     (Mueller and Stickney,  1970). The change in structure promotes the characteristic cracking of
21     stressed/stretched rubber called "weathering." A 5% tensile strain will produce cracks on the
22     surface of the rubber that increase in number with increased stress/stretching. The rate of crack
23     growth is dependent on the degree of stress, the type of rubber compound, ozone concentration,
24     time of exposure, ozone velocity, and temperature (Bradley and Haagen-Smit,  1951; Lake and
25     Mente, 1992) (Gent and McGrath, 1965). Once cracked, there is further ozone penetration and
26     in additional cracking and eventually mechanical weakening or stress relaxation (U.S.
27     Environmental Protection Agency, 1996). Razumovskii et al. (1988) demonstrated the effect of
28     ozone on stress relaxation of polyisoprene vulcanizates. A decrease in stress (stress relaxation)
29     was caused by ozone-induced cracks in exposed elastomers resulting in irreversible changes in
30     the elastomer dimensions and decreased tensile strength. A description of findings of earlier
       January 2005                             11-1        DRAFT-DO NOT QUOTE OR CITE

-------
 1      studies appear in the 1996 Air Quality Criteria Document for ozone (U.S. Environmental
 2      Protection Agency, 1996).
 3           To counteract the effect of ozone on elastomers, antiozonants and wax have been added to
 4      the elastomeric formulations during processing. An antiozonant is an additive used to protect a
 5      polymer against the effects of ozone-induced degradation and, hence, used mainly in diene
 6      rubbers.  Antiozonant protection works either by providing a physical barrier to ozone
 7      penetration forming a thin surface film of an ozone-resisting wax or by chemically reacting with
 8      ozone or polymer ozonolysis products, as do aromatic diamines such as p-phenylene diamine
 9      derivatives.  The antiozonant diffuses to the surface of the elastomeric material where it reacts
10      with ozone faster than ozone reacts to break the molecular chain and the carbon-carbon double
11      bond, or the antiozonant diffuses to the surface of the material but is not reactive with ozone and
12      serves as a protective coating against ozone attack. The antiozonant may also serve to  scavenge
13      ozone while also providing protective film against ozone attack (Andries et al., 1979; Lattimer
14      etal., 1984).
15           Most of the studies on ozone effects on elastomers were designed to evaluate the
16      effectiveness of antiozonants in counteracting the rubber cracking produced by ozone exposure.
17      Consequently, many of the studies were conducted using ozone concentrations higher than those
18      typically found in the ambient air. Natural rubber strips exposed to high concentrations of ozone
19      (20,000 ppm) under stressed conditions cracked almost instantaneously and were broken within
20      1 sec.  When the ozone concentration was lowered (0.02 to 0.46 ppm), the time to  required to
21      produce cracks in the exposed rubber material  was increased (Bradley and Haagen-Smit, 1951).
22      Lake and Mente (1992) studied the effect of temperature on ozone-induced  elastomer cracking
23      and antiozonant protection on natural rubber, epoxidised natural rubber, and two acrylonitrile-
24      butadiene copolymers under constant strain. Temperatures ranged from -20° C to +70° C. The
25      elastomers were exposed to 0.05 to 1,000 ppm ozone for 16 h. Ozone cracking decreased at
26      lower ambient temperatures, however, diffusing of both chemical and wax antiozonants also
27      were slowed  at the lower temperatures.  Cracking was slightly increased at the higher
28      temperatures but the antiozonants  offered more protection.
29           Serrano et al. (1993) evaluated the appropriateness of using ozone-induced elastomer
30      cracking to estimated the ambient  ozone concentrations. Two vulcanized natural rubber
31      compounds were exposed for 24 h to varying ozone concentrations under stressed  conditions.

        January 2005                              11 -2        DRAFT-DO NOT QUOTE OR CITE

-------
 1
 2
 3
 4
 5
 6
 7
Ozone concentrations were 60, 80, 90, 100, and 120 ppb for durations of 2, 4, or 6 h. The 24 h
average ozone concentrations ranged from 31 to 57.5 ppb. There was a clear relationship
between the 24-h average ozone concentration and the distribution of crack length frequencies
on the rubber surface.  Table 11-1 gives the average 24-h ozone concentration and lengths for
two vulcanized natural rubber strips.
          Table 11-1. Average 24-h Ozone Concentrations Producing the Highest Frequency of
          Cracks of a Certain Length in the Middle and Central Zones of the Rubber Test Strips
1% Antiozonant 4010NA #
Crack Length (mm)
0.05-0.10
0.10-0.15
0.15-0.20
0.20 - 0.40
Middle Zones
37.5
45.0
48.0
> 57.5
Central Zones
37.5
48.0
> 57.5
> 57.5
0.5% Antiozonant 4010NA
Middle Zones
40.0
48.0
> 57.5
> 57.5
Central Zones
42.5
53.0
> 57.5
> 57.5
        Ozone concentrations given in ppb.
        Adapted from Serrano et al. (1993).
 1      11.11.2  Textiles and Fabrics
 2           Ozone can damage textiles and fabrics by methods similar to those associated with
 3      elastomers.  Generally, synthetic fibers are less affected by ozone than natural fibers, however,
 4      ozone contribution to the degradation of textiles and fabrics is not considered significant (U.S.
 5      Environmental Protection Agency, 1996). A study reported by Bogaty et al. (1952) showed that
 6      ozone effects moistened cloth more than dry cloth. Scoured cotton duck cloth and commercially
 7      bleached cotton print cloth were exposed to 20 to 60 ppb for 1,200 h (50 days). The rate of
 8      deterioration was measured by the changes in cuprammonium fluidity values and the fabric
 9      breaking strength.  At the end of the 1,200-h exposure, there was a 20% loss in breaking
10      strength. Table 11-2 list the changes in cuprammonium fluidity values for both fabrics.
       January 2005
                                                     DRAFT-DO NOT QUOTE OR CITE

-------
                 Table 11-2. Cuprammonium Fluidity of Moist Cotton Cloth Exposed
                                        to 20 to 60 ppb Ozone
                                  Duration of Exposure (h)    Cuprammonium Fluidity (rhes)
        Duck Cloth                          0                             2.6
                                            200                            2.8
                                            680                            4.0
                                            960                            6.8
                                            1200                           9.5
        Bleached Print Cloth                  0                             8.2
                                            200                            8.7
                                            510                            9.4
                                            650                           12.0
                                            865                           12.7
                                            1500                          16.5

        Adapted from Bogaty et al. (1952).
 1     11.11.3  Dyes, Pigments, and Inks
 2          Ozone fading of textile dyes is diffusion-controlled; the rate of fading is controlled by the
 3     diffusion of the dye to the fiber surface. Many textile dyes react with ozone; however, the rate
 4     and severity of the ozone attack is influenced by the chemical nature of the textile fiber and the
 5     manner in which the dye is applied.  Ozone molecules break the aromatic ring portion of the dye
 6     molecule, oxidizing the dye (U.S. Environmental Protection Agency, 1996).  In case of aromatic
 7     azo dyes, ozone attacks the aromatic rings and electron rich nitrogen atoms (Matsui et al., 1988).
 8     Grosjean et al. (1987; 1988a,b) proposed a mechanism of reaction of ozone and indigo,
 9     thioindigo, and dibromoindigo, alazarin, and curcumin dyes under dark conditions.  Ozone
10     attaches to the dye molecule at the unsaturated carbon = carbon bond.  An ozone adduct is
11     formed (1,2,3-trioxolane), followed by scission of the carbon-carbon bond and the subsequent
12     formation of the corresponding Criegee biradical. A similar mechanism was proposed for the
13     reaction of ozone with triphenylmethane colorant Basic Violet 14.  Ozone attacked Basic Violet
14     14 at the carbon=carbon unsaturated bond and at the carbon-nitrogen unsaturated bond under
15     dark conditions.  Other members of the group of triphenylmethane colorants with unsaturated
16     carbon-carbon bonds also are expected to be subject to ozone fading.  Tripheylmethane
17     colorants that are expected to be ozone-fugitive include the amino-substituted cationic dyes

       January 2005                             11 -4        DRAFT-DO NOT QUOTE OR CITE

-------
 1      (Malachite Green, Brilliant Green, Crystal Violet, Pararosaniline Chloride, Methyl Green, and
 2      others) (Grosjean et al., 1989).
 3           An indication that ozone caused textile dye fading was first reported by Salvin and Walker
 4      (1955). The researchers found that the fading was primarily the result of the destruction of the
 5      blue dye molecule. Drapes made of acetate, Arnel, and Dacron and dyed with anthraquinone
 6      blue dye exhibited a decrease in shade that was not accompanied by the characteristic reddening
 7      caused by NOX. Figures 11-1 and 11-2 demonstrate the effect of ozone exposure on nylon 6 yarn
 8      colored with several blue dyes. Nylon samples inside the home were located on a wall away
 9      from sunlight. Outside nylon samples were placed  on a covered patio or under the eaves of the
10      house to minimize exposure to sunlight and rain. Ozone concentrations ranged from 2 to 5 ppb
11      outside and 0 to 2 ppb inside. The percent change in dye color was determined monthly by
12      extraction and analysis of the remaining dye or by instrumental measurement of the color change
13      (Haylock and Rush, 1978).
14
15      11.11.4  Artists' Pigments
16           Several artists' pigments are sensitive to fading and oxidation by ozone when exposed to
17      concentrations found in urban areas (Shaver et al., 1983; Drisko et al., 1985; Whitmore et al.,
18      1987; Whitmore and Cass, 1988; Grosjean et al., 1993).  The organic pigments that are ozone
19      fugitive include alizarin red pigments containing lakes of the poly cyclic aromatic compound
20      1,2-dihydroxyanthraquinone, blue-violet pigments containing substituted triphenylmethane
21      lakes, indigo, and yellow coloring agents containing polyfunctional, polyunsaturated compounds
22      such as curcumin (Grosjean et al., 1987). Because of the potential of ozone to damage works of
23      art, recommended limits on ozone concentrations in museums, libraries, and archives are
24      relatively low, ranging from 0.013 to 0.01 ppm.
25           Experimental studies demonstrate a concentration x time (C x T) relationship between
26      ozone concentration and exposure time and pigment fading.  Cass et al. (1991) summarized
27      some of the earlier research on the effects of ozone on artists' pigments. In studies evaluating
28      the effect of ozone on organic and inorganic watercolors and traditional organic pigments only
29      the traditional organic pigments showed measurable fading from ozone exposure.  Of the
30      inorganic pigments tested, only the arsenic sulfides showed ozone-related changes. The
        pigments were exposed to 0.3 to 0.4 ppm ozone for 3 mo in the absence of light, at 22 °C and

        January 2005                              11-5        DRAFT-DO NOT QUOTE OR CITE

-------
                                    100 r
                                      J  FMAMJ  J ASOND J
                                                   Months
       Figure 11-1.  In-service fading of nylon 6 yarn inside house.  • = C.I. Disperse Blue 3;
                     O = C.I. Basic Blue 22; A = C.I. Acid Blue 27; x = C.I. Disperse Blue 56;
                     A = C.I. Acid Blue 232.
       Source: Haylock and Rush (1978).
 1     50% RH.  The authors equated this exposure to a C x T of 6 to 8 years inside a Los Angeles
 2     museum with air conditioning but without a pollutant removal system.
 3          Whitmore and Cass (1988) studied the effect of ozone on traditional Japanese colorants.
 4     Most of these compounds are insoluble metal salts that are stable in light and air.  Suspensions or
 5     solutions of the colorants were airbrushed on hot-pressed watercolor paper or silk cloths.
 6     A sample of Japanese woodblock print also was included in the analysis.  Samples were exposed
 7     to 0.4 ppm ozone at 22 °C, 50% relative humidity, in the absence of light for 12 wk.  Changes in
 8     reflectance spectra were used to evaluate the effect of ozone exposure on colorant fading.
 9     Among the colorants tested on paper, curmin, indigo, madder lake, and lac lake were the most
10     sensitive to ozone exposure.  Gamboge was relatively insensitive to ozone.
       January 2005
11-6
DRAFT-DO NOT QUOTE OR CITE

-------
                                    100 r
                                     90 -
                                       J FMAMJ  J  ASOND J
                                                    Months
       Figure 11-2.  In-service fading of nylon 6 yarn outside house.  • = C.I. Disperse Blue 3;
                     O = C.I. Basic Blue 22; A = C.I. Acid Blue 27; x = C.I. Disperse Blue 56;
                     A = C.I. Acid Blue 232.
       Source: Haylock and Rush (1978).
 1     The blue and green areas of the sample from the woodblock print was very reactive due to the
 2     indigo dye ozone sensitivity.  The other colorants, red, yellow, and purple, showed very little
 3     sensitivity to ozone.  The textiles dyes that reacted with ozone were indigo, alone or in
 4     combination with several yellow dyes.
 5           Ye et al. (2000) reported the rate of ozone fading of traditional Chinese plant dyes.  Twelve
 6     different colorants were applied to watercolor paper and silk and exposed to 0.4 ppm ozone at
 7     25° C, at 50% RH, in the absence of light for 22 wks.  Dye fading was greater when the colorant
 8     was applied to the watercolor paper compared to the silk cloth due to the darker initial depth of
 9     the shade, the greater saturation of the colorant throughout the cloth. Turn eric, gromwell, and
10     violet on paper was particularly reactive. Tangerine peel was moderately reactive and sappan
11     wood, dalbergia wood, Chinese gall, indigo, and Chinese yellow cork tree were slightly reactive
       January 2005
11-7
DRAFT-DO NOT QUOTE OR CITE

-------
 1      to ozone. Black tea was not reactive to ozone. The colorants on silk samples showing color
 2      changes were gromwell, sappan wood, gardenia, tummeric, and violet.  Figures 11-3 and 11-4
 3      demonstrate the color change of the various colorants on watercolor paper and silk.
 4           Artists' pigments also have exhibited fading when exposed to a mixture of photochemical
 5      oxidants. Grosjean et al. (1993) exposed 35 artists' pigments to a mixture of photochemical
 6      oxidants consisting of ozone, nitrogen dioxide (NO2), and peroxyacetyl nitrate (PAN) for
 7      12 wks.  Weekly average photochemical concentrations were 200 ppb for ozone, 56 ± 12 to
 8      99 ± 24 for NO2, and Il±3tol8±2 for PAN. All exposures were carried out at room
 9      temperature in the absence of light. To determine the effect of humidity on pigment fading, the
10      relative humidity was increased from 46% after 8 weeks of exposure to 83% for a 2 week period
11      and then returned to 46% for the remainder of the exposure.
12           Table 11-3 lists the artists' pigment and degree of fading. Eleven of the pigments
13      exhibited negligible color change, 12 had small color changes, 3 had modest color changes, and
14      9 exhibited substantial color changes. Fading of Disperse Blue 3 and Reactive Blue 2 were
15      likely the result of NO2 exposure, the fading of triphenylmethanes is consistent with exposure to
16      nitric acid formed under high humidity conditions. Fading of the indigos was dominated by
17      ozone exposure and curcumin was faded by all of the photochemicals studied. Increasing the
18      relative humidity resulted in a substantial color change  for all of the pigments, with the
19      exception of curcumin and indigo.
20
21      11.11.5  Surface Coatings
22           Ozone will act to erode some surface coatings (paints, varnishes, and lacquers).  However,
23      many of the available studies on ozone degradation of surface coatings  do not separate the
24      effects of ozone from  other pollutants or environmental factors such as  weather, humidity, and
25      temperature.  Campbell et al. (1974) attempted to demonstrate an ozone related effect on oil
26      house paint, acrylic latex coating, alkyd industrial maintenance coating, urea alkyd coil coating,
27      and nitrocellulose/acrylic automotive paint. Painted test panels were exposed to 100 and
28      1,000 ppb ozone in a xenon arc accelerated weathering  chamber for up  to 1,000 h.  Using weight
29      loss as a measure of ozone-induced erosion the researchers concluded that all of the paints tested
30      suffered degradation in the presence of ozone and that the automotive finish suffered the most
        January 2005                              11-8        DRAFT-DO NOT QUOTE OR CITE

-------
                                 35
                                 30
                                 25
                              0)
                              o  20
                                 15
                              o
                              o
                              o
                                 10
                                         zi cao (gromwe
                                         zi ding cao (violet)
                                           5       10       15       20
                                     Weeks of Exposure to 0.40 ppm Ozone
                                           very reactive colorants
                               e 3
                               o
                               o
                               o
                                    -0- su mu (sappan wood)
                                    -*- Jiang xiang (dalbergia wood)
                                    -Ar hong chaye (black tea)
                                   . -Oh wu bei zi (Chinese gall)
                                    -O- ban Ian gen (indigo)
                                    "V~ ju zi pi (tangerine peel)
                                    "A- huang bai (Chinese yellow cork tree)
                                  0       5       10       15      20
                                    Weeks of Exposure to 0.40 ppm Ozone
                                        moderately reactive colorants
Figure 11-3.   Observed color changes for natural colorant-on-paper systems during
                exposure to 0.40 ppm ozone at 25 °C ± 1 °C, 50% RH, in the absence of light.

Source:  Ye et al. (2000).
January 2005
11-9
DRAFT-DO NOT QUOTE OR CITE

-------
                               a)
                               1
                               D
                               o 2
                               o
                               O
                                        . su mu (sappan wood)
                                        - zi cao (gromwell)
                                         zi ding cao (violet)
                                        - Jiang huang (tumeric)
                                        - huang zhi zi (gardenia)
                                                                     J	=
                                                   10
                                                            15
                                                                    20
                                     Weeks of Exposure to 0.40 ppm Ozone
                                             reactive colorants
                                 2.0
                                 1.5
                              HI
                              
-------
                 Table 11-3. Color Change After 12 Weeks of Exposure to a
                               Mixture of Photochemical Oxidants
 Colorant*
Color Change (A£" units)f
 Chemical Functionality or
 Chemical Composition
 Acid Red 37 (17045)*


 Acid Yellow 65 *


 Alizarin Carmine

 Alizarin Crimson
 (Pigment Red 83)

 Aurora Yellow (77199)

 Basic Fuschin (42510)*

 Brilliant Green (42040)*

 Brown Madder

 Cadmium Yellow (77199)

 Carmine


 Chrome Yellow (77600)*

 Copper phthalocyamne
 (Pigment Blue 15)

 Crimson Lake

 Curcumin (Natural Yellow 3)


 Disperse Blue 3

 French Ultramarine Blue

 Gamboge (Natural Yellow 24)

 Hooker's Green Light


 Indigo (a formulation)


 Indigo carmine *

 Indigo (73000)*

 Mauve


 New Gamboge
       11.7 ±0.5


        1.8 ±0.5


        1.8 ±0.2

        1.4 ±0.2


        0.5 ±0.1

       33.4 ±3.0

       20.6 ±2.1

        1.7±0.1

        0.4 ±0.1

        1.8 ±0.2


        1.7 ± 1.2

        l.OiO.l


        3.5 ±0.3

       15.2 ±2.6


       10.8 ±0.1

        0.8 ±0.3

        0.4 ±0.1

        1.5 ±0.4


        l.liO.l


       14.0 ±1.9

       64.1 ±4.5

        3.6 ±0.5


        0.9 ±0.1
 Aminophenyl-substituted azo dye,
 sulfonate salt

 Nitro- and phenyl-substituted azo dye,
 sulfonate salt

 Alizarin lake

 Alizarin lake
 Cadmium sulfide

 Amino-substituted triphenylmethane

 Amino-substituted triphenylmethane

 Alizarin lake

 Cadmium sulfide

 Lake of cochineal (substituted
 anthraquinone)

 Lead chromate

 Copper phthalocyanine


 Alizarin lake

 1,7 bis (4-hydroxy-3-methoxyphenyl)-
 1,6-heptadiene-3,5-dione

 Amino-substituted anthraquinone


 Gambogic acid

 Chlorinated copper phthalocyanine plus
 ferrous beta naphthol derivative

 Alizarin lake plus lampblack plus copper
 plthalocyanine

 5,5-indigo disulfonic acid, sodium salt
 Lake of triphenyl methane (basic fuschin)
 plus copper phthalyocyanine

 Arylamide yellow (CI11680) plus
 toluidine red
January 2005
             11-11
DRAFT-DO NOT QUOTE OR CITE

-------
                   Table 11-3 (cont'd).  Color Change After 12 Weeks of Exposure to a
                                    Mixture of Photochemical Oxidants
        Colorant*
Color Change (A£" units)f
 Chemical Functionality or
 Chemical Composition
        Pararosaniline base (42500)*

        Payne's Grey


        Permanent Magenta

        Permanent Rose

        Prussian Blue


        Prussian Green

        Purple Lake

        Reactive Blue 2 (61211)*


        Rose Carthane (12467)


        Rose Dore

        Thioindigo Violet (73312)*

        Winsor Yellow (11680)
       25.6 ±4.7

        1.0 ±0.1


        l.liO.l

        2.0 ±0.1

        0.7 ±0.2
        1.6 ±0.3

        0.9 ±0.2

        2.3 ±0.3

       14.4 ± 1.1


        0.8 ±0.2


        2.0 ±0.2

        1.9 ±1.2

        0.5 ±0.2
 Amino-substituted triphenylmethane

 Alizarin lake plus prussian blue plus
 lampblack plus ultramarine blue

 Quinacridone

 Quinacridone

 Ferric ferrocyanide


 Arylamide yellow plus prussian blue

 Alizarin lake

 Amino-substituted anthraquinone,
 sulfonate salt

 Arylamide (Pigment Red 10) plus
 xanthene (Pigment Red 90)

 Quinacridone plus Yellow 3

 Chlorinated thioindigo

 Arylamide yellow
        * On watercolor paper unless otherwise indicated. Color Index (CI) names or CI numbers are given in
          parentheses.
        •f Mean ± one standard deviation for triplicate samples calculated from the parameters L*, a*, and b* measured
          with the color analyzer.
        { On Whatman 41 paper.

        Source:  Grosjeanetal. (1993)
1      ozone-induce degradation. When ozone degradation was measured using scanning electron

2      microscopy, the oil house paint and latex coating samples showed erosion above that seen with

3      clean air but only at the highest exposure level. No effects were noted for the automotive paint.

4      The other painted surfaces were not evaluated.

5            Spence et al. (1975) studied the effect of air pollutants and relative humidity on oil based

6      house paint, acrylic latex house paint, acrylic coil coating, and vinyl coil coating under

7      laboratory conditions.  Test panels were exposed in weathering chambers equipped with a xenon

8      are light for simulating sunlight to low and high levels of ozone (0.08 and 0.5 ppm), sulfur
       January 2005
            11-12
DRAFT-DO NOT QUOTE OR CITE

-------
 1     dioxide (0.03 and 0.5 ppm), and nitrogen dioxide (0.05 and 0.5 ppm) and relative humidity
 2     (50 and 90%). Samples were exposed for a total of 1000 h. The exposure cycle consisted of
 3     20 min of dew and 20 min of light. The effects of the exposure on the painted surfaces were
 4     measured by weight loss and loss in film thickness. The acrylic coil coating had the lowest
 5     erosion rate of the surface coatings tested.  However, ozone was the only pollutant that had a
 6     significant effect on the surface erosion. Sulfur dioxide and relative humidity were significant
 7     factors in the erosion of oil base house paints and vinyl coil coating. The findings for acrylic
 8     latex house paint were not reported.
 9
10
       January 2005                              11-13       DRAFT-DO NOT QUOTE OR CITE

-------
  1       REFERENCES
  2
  3       Andries, J. C.; Rhee, C. K.; Smith, R. W.; Ross, D. B.; Diem, H. E. (1979) A surface study of ozone attack and
  4             antiozonant protection of carbon black loaded natural rubber compounds. Rubber Chem. Technol.
  5             52: 823-837.
  6       Bogaty, H.; Campbell, K. S.; Appel, W. D. (1952) The oxidation of cellulose by ozone in small concentrations.
  7             Text. Res. J. 22: 81-83.
  8       Bradley, C. E.; Haagen-Smit, A. J. (1951) The application of rubber in the quantitative determination of ozone.
  9             Rubber Chem. Technol. 24: 750-755.
10       Campbell, G. G.; Schurr, G. G.; Slawikowski, D. E.; Spence, J. W. (1974) Assessing air pollution damage to
11             coatings. J. Paint Technol. 46: 59-71.
12       Cass, G. R. (1991) Deterioration of materials due to ozone exposure: current problems and future research needs. In:
13             Berglund, R. L.; Lawson, D. R.; McKee, D. J., eds. Tropospheric ozone and the environment: papers from an
14             international conference; March 1990; Los Angeles, CA. Pittsburgh, PA:  Air & Waste Management
15             Association; pp. 311-320. (A&WMA transactions series no. TR-19).
16       Drisko, K.; Cass, G. R.; Whitmore, P. M.; Druzik, J. R. (1985) Fading of artists' pigments due to atmospheric ozone.
17             In: Vendl, A.; Pichler, B.; Weber, J.; Banik, G., eds. Wiener Berichte tiber Naturwissenschaft in der Kunst:
18             Doppelband 2/3. Vienna, Austria: Verlag ORAC; pp. 66-87.
19       Gent, A. N.; McGrath, J. E. (1965) Effect of temperature on ozone cracking of rubbers. J. Polymer Sci. A
20             3:  1473-1482.
21       Grosjean, D.; Whitmore, P. M.; De Moor, C. P.; Cass, G. R.; Druzik, J. R. (1987) Fading of alizarin and related
22             artists' pigments by atmospheric ozone: reaction products and mechanisms. Environ. Sci. Technol.
23             21:635-643.
24       Grosjean, D.; Whitmore, P. M.; Cass, G. R.; Druzik, J. R. (1988a) Ozone fading of natural organic colorants:
25             mechanisms and products of the reaction of ozone with indigos. Environ.  Sci. Technol. 22: 292-298.
26       Grosjean, D.; Whitmore, P. M.; De Moor, C. P.; Cass, G. R.; Druzik, J. R. (1988b) Ozone fading of organic
27             colorants: products and mechanism of the reaction of ozone with curcumin. Environ. Sci. Technol.
28             22: 1357-1361.
29       Grosjean, D.; Whitmore, P. M.; Cass, G. R.; Druzik, J. R. (1989) Ozone fading of triphenylmethane colorants:
30             reaction products and mechanisms. Environ. Sci. Technol. 23: 1164-1167.
31       Grosjean, D.; Grosjean, E.; Williams, E. L., II.  (1993) Fading of artists' colorants by a mixture of photochemical
32             oxidants. Atmos. Environ. Part A 27: 765-772.
33       Haylock, J. C.; Rush, J. L. (1978) Studies on the ozone fading of anthraquinone  dyes on nylon fibers. Part II:
34             In-service performance. Text. Res. J. 48: 143-149.
35       Lake, G. J.; Mente, P. G. (1992) Ozone cracking and protection of elastomers at high and low temperatures. J. Nat.
36             Rubber Res. 7: 1-13.
37       Lattimer, R. P.; Layer, R. W.; Rhee, C. K. (1984) Mechanisms of antiozonant protection: antiozonant-rubber
3 8             reactions during ozone exposure. In: 32nd annual conference on mass spectrometry and allied topics.
39             May/June; San Antonio, TX. Bethesda, MD: American Society for Mass  Spectrometry; pp. 357-358.
40       Matsui, M.; Koike, T.; Shibata, K. (1988) Ozone fading of phenolphthalein and aurin. J. Soc. Dyers Colour
41             104:482-486.
42       Mueller, W. J.; Stickney, P. B. (1970) A survey and economic assessment of the effects of air pollution on
43             elastomers:  final report. Columbus, OH: Battelle Memorial Institute, Columbus Laboratories; National Air
44             Pollution Control Administration contract no. CPA-22-69-146.
45       Razumovskii, S. D.; Podmasteriev,  V. V.; Zaikov, G. E.  (1988) Kinetics and mechanism of stress relaxation of
46             polyisoprene vulcanizates under ozone ageing. Polym. Degrad. Stab. 20: 37-47.
47       Salvin, V. S.; Walker, R. A. (1955) Service fading of disperse dyestuffs by chemical agents other than the oxides of
48             nitrogen. Text. Res. J. 25: 571-585.
49       Serrano, E.; Castro, M.; Macias, A. (1993) An improved direct method of rubber cracking analysis for estimating
50             24-hour ozone levels. Atmos. Environ. Part A 27: 431-442.
51       Shaver, C. L.; Cass, G. R.; Druzik, J. R. (1983) Ozone and the deterioration of works of art. Environ. Sci. Technol.
52             17:748-752.
53       Spence, J. W.; Haynie, F.; Upham, J. B. (1975) Effects of gaseous pollutants on paints: a chamber study. J. Paint
54             Technol. 47: 57-63.
         January 2005                                  11-14         DRAFT-DO NOT QUOTE OR CITE

-------
 1       U.S. Environmental Protection Agency. (1996) Air quality criteria for ozone and related photochemical oxidants.
 2             Research Triangle Park, NC: Office of Research and Development; report nos. EPA/600/AP-93/004aF-cF. 3v.
 3             Available from: NTIS, Springfield, VA; PB96-185582, PB96-185590, and PB96-185608. Available:
 4             http://cfpub2.epa.gov/ncea/.
 5       Whitmore, P. M; Cass, G. R. (1988) The ozone fading of traditional Japanese colorants. Stud. Conserv. 33: 29-40.
 6       Whitmore, P. M.; Cass, G. R.; Druzik, J. R. (1987) The ozone fading of traditional natural organic colorants on
 7             paper. J. Am. Inst. Conserv. 26: 45-58.
 8       Ye, Y.; Salmon, L. G.; Cass, G. R. (2000) The ozone fading of traditional Chinese plant dyes. J. Am. Inst. Conserv.
 9             39:245-257.
10
         January 2005                                   11-15        DRAFT-DO NOT QUOTE OR CITE

-------